Document 11386030

advertisement
 Genetic Analysis of Synaptogyrin Function in the Synaptic Vesicle Cycle by Robin J. Stevens B.S. Biological Sciences Stanford University, 2005 SUBMITTED TO THE DEPARTMENT OF BIOLOGY IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY IN BIOLOGY AT THE MASSACHUSETTS INSTITUTE OF TECHNOLOGY FEBRUARY 2012 © 2012 Robin J. Stevens. All Rights Reserved. The author hereby grants to MIT permission to reproduce and to distribute publicly paper and electronic copies of this thesis document in whole or in part in any medium now known or hereafter created. Signature of author: _____________________________________________________________________________ Department of Biology November 28, 2011 Certified by: _____________________________________________________________________________________ Dr. J. Troy Littleton Professor of Biology Thesis Supervisor Accepted by: _____________________________________________________________________________________ Dr. Alan D. Grossman Praecis Professor of Biology Co-­‐Chairman of the Graduate Committee 2 Genetic Analysis of Synaptogyrin Function in the Synaptic Vesicle Cycle by Robin J. Stevens Submitted to the Department of Biology on January 17, 2012 in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in Biology ABSTRACT Neuronal communication relies on the continual replenishment of synaptic vesicles that are primed for neurotransmitter release in response to action potentials. A vast array of proteins is required to mediate synaptic vesicle biogenesis, trafficking, docking, exocytosis, and endocytosis. Synaptogyrin and synaptophysin are abundant and evolutionarily conserved synaptic vesicles proteins that were identified over twenty years ago, yet their exact function in the synaptic vesicle cycle remains unknown. To further elucidate the role of these proteins, we have generated and characterized a synaptogyrin null mutant in Drosophila, whose genome encodes a single synaptogyrin isoform and lacks a synaptophysin homolog. Here we demonstrate that Drosophila synaptogyrin is abundantly expressed in neurons, where it localizes to the presynaptic terminal of the larval neuromuscular junction (NMJ). Drosophila lacking synaptogyrin are viable and fertile and have no overt deficits in motor function or courtship behavior. Ultrastructural analysis of mutant larvae revealed an increase in average synaptic vesicle diameter as well as enhanced variability in the size of synaptic vesicles. In addition, the resolution of endocytic cisternae into synaptic vesicles in response to robust exocytosis is defective in synaptogyrin mutants. While basal synaptic transmission at the larval NMJ is unaffected, synaptogyrin mutants do display increased facilitation during high-­‐frequency stimulation, indicating that synaptic vesicle exocytosis is abnormally regulated during strong stimulation conditions. These results suggest that, while not required for neurotransmission, Drosophila synaptogyrin nevertheless modulates synaptic vesicle exo-­‐endocytosis, especially during elevated rates of synaptic vesicle fusion. Thesis Supervisor: Dr. J. Troy Littleton Title: Professor of Biology 3 4 Acknowledgments I would like to start by thanking my advisor, Troy Littleton, for his support and guidance over the years as well as for giving me the opportunity to work in such a wonderful laboratory. The Littleton Lab is not just a scientifically stimulating environment, it also is a place filled with fantastic people who make each day enjoyable, even if the science isn’t going the way you’d hoped. Thanks to all members of the Littleton Lab, past and present, especially Yulia Akbergenova and Ramon Jorquera for their invaluable contributions to the synaptogyrin project, as well as Avi Rodal and Joost Schulte for their endless knowledge and assistance, and Dina Volfson and Charles Moss for making sure the lab runs smoothly. I would also like to thank my MSRP student, Joe Endicott, who was a huge help in the short time he was a member of our lab. I also would like to thank Dennis Kim, Martha Constantine-­‐Paton, and Weifeng Xu for serving on my committee and for their helpful comments and suggestions. I could not have gotten through all the ups and downs of graduate school without the support of my friends here at MIT/Cambridge and the rest of the Biology entering class of 2005. I would especially like to thank Jen, Jan, Cindy, Aline, Javier, Lauren, and Kurt for encouraging me to have a life outside of lab. Thanks to my extended family for their love and encouragement, especially my maternal grandparents, Jean and Ty, who never cease to tell me how proud they are of me. I also have to thank my brother Brian for being the most loving, caring, and encouraging brother I could ever imagine having. I am so lucky to have a brother I can call a best friend. Finally, I would like to thank my parents for supporting me every step of the way on this long journey. Mom, you helped shape me into the woman I am today. You instilled in me the importance of education, the value of standing up for what is right, and the desire to reach out and help others. Dad, I simply would not be a scientist if it weren’t for you. You not only helped me with science fair projects and took me on field trips to your lab, you also showed me how fun science can be. Growing up I wanted to be just like you – a scientist with a Ph.D. Although it has been difficult at times, because of your support there was never a moment I doubted I could achieve that goal. 5 6 Table of Contents Chapter 1: Synaptogyrin, Synaptophysin, and the Synaptic Vesicle Cycle ................... 9 Introduction .................................................................................................................................................10 The discovery of synaptic vesicles .....................................................................................................10 The synaptic vesicle cycle ......................................................................................................................12 Clathrin-­‐mediated vs. kiss-­‐and-­‐run endocytosis .........................................................................14 Clathrin-­‐mediated endocytosis ...........................................................................................................16 Bulk endocytosis ........................................................................................................................................19 Synaptic vesicle pools..............................................................................................................................21 The SNARE complex and regulated synaptic vesicle fusion ...................................................23 Synaptotagmin and complexin ............................................................................................................24 Molecular anatomy of synaptic vesicles ..........................................................................................27 Initial characterization of synaptophysin and synaptogyrin..................................................28 Synaptophysin ........................................................................................................................................28 Synaptogyrin...........................................................................................................................................29 Physins and gyrins: evolutionarily related MARVEL domain proteins.............................30 Synaptoporin, pantophysin, and mitsugumin 29 ........................................................................33 Synaptogyrins 2-­‐4 .....................................................................................................................................34 SCAMPs ..........................................................................................................................................................34 Potential functions of synaptophysin and synaptogyrin..........................................................36 Regulation of exocytosis .....................................................................................................................36 Regulation of endocytosis ..................................................................................................................41 Synaptophysin and synaptogyrin knockouts ................................................................................43 Synaptogyrin: a schizophrenia susceptibility gene? .................................................................47 The Drosophila larval NMJ as a model of synaptic development and function...............48 References ....................................................................................................................................................49 Figures............................................................................................................................................................67 Chapter 2: Characterization of the Drosophila Synaptogyrin Homolog ......................69 Introduction .................................................................................................................................................70 Results ............................................................................................................................................................74 Evolutionary analysis of synaptogyrin and synaptophysin...................................................74 Drosophila synaptogyrin is present in neurons and targeted to synaptic vesicles......77 Generation of a Drosophila synaptogyrin mutant ..................................................................78 Behavioral analysis of gyr mutants reveals no significant defects in climbing ability or courtship..........................................................................................................................81 Discussion .....................................................................................................................................................84 Drosophila synaptogyrin is a synaptic vesicle protein ...........................................................84 The loss of synaptogyrin does not significantly impact certain adult behaviors..........85 Synaptogyrin is not required for viability, fertility, or basic motor function.................86 Methods .........................................................................................................................................................88 References ....................................................................................................................................................97 Figures.........................................................................................................................................................102 7 Chapter 3: Analysis of Synaptogyrin’s Role in Synaptic Vesicle Exo-­endocytosis121 Introduction ..............................................................................................................................................122 Results .........................................................................................................................................................125 gyr mutants have normal synaptic growth and bouton morphology ............................125 Ultrastructural analysis of gyr larvae reveals alterations in synaptic vesicle diameter and density ...................................................................................................................127 gyr mutants have an increased number of endocytic cisternae following intense stimulation.......................................................................................................................................129 Synaptic vesicle diameter in gyr mutants shifts to more normal values following a high K+ shock ...............................................................................................................................132 gyr mutants exhibit increased release probability and facilitation................................133 Discussion ..................................................................................................................................................136 The relationship between synaptogyrin and synaptic activity .........................................137 A preliminary model to explain synaptogyrin’s role in synaptic vesicle formation ..138 Synaptogyrin influences synaptic vesicle release probability............................................ 140 Methods.......................................................................................................................................................143 References..................................................................................................................................................148 Figures .........................................................................................................................................................152 Chapter 4: Conclusions and Perspectives...................................................................................169 Summary.....................................................................................................................................................170 The loss of synaptogyrin and synaptophysin results in relatively mild phenotypes ...........................................................................................................................................170 Phylogenetic analysis of synaptophysin and synaptogyrin..................................................172 Structure-­‐function analysis of synaptogyrin and synaptophysin......................................174 The role of synaptogyrin and synaptophysin in endocytosis ..............................................176 Further analysis of synaptogyrin’s role in vesicle biogenesis and endocytosis ..........178 Examination of genetic interactions with Drosophila synaptogyrin ................................180 Conclusions................................................................................................................................................184 References..................................................................................................................................................185 8 Chapter 1 Synaptogyrin, Synaptophysin, and The Synaptic Vesicle Cycle Robin J. Stevens The Picower Institute for Learning and Memory, Department of Biology, Massachusetts Institute of Technology, Cambridge, MA 02139 9 Introduction One of the early debates in the field of neurobiology centered on the nature of synaptic transmission, namely whether communication between neurons was conducted via chemical or electrical means (reviewed in (Eccles, 1982)). Key discoveries, including the identification of a synaptic delay and the detection of synaptic inhibition using intracellular recordings, eventually demonstrated that chemical transmission occurs at a majority of synapses, while gap junction-­‐mediated electrical transmission is less common (Kuffler, 1942b, a; Brock et al., 1952). It has since been established that chemical neurotransmitters are stored in synaptic vesicles and are released in discrete packets upon calcium entry into the presynaptic terminal. Although small in size and simple in appearance on electron micrographs, synaptic vesicles are complex organelles containing dozens of proteins that are responsible for the targeting, fusing, and recycling of these vesicles (reviewed in (Sudhof, 2004)). Among these proteins are synaptophysin and synaptogyrin, two evolutionarily related proteins that were identified as synaptic vesicle components over twenty years ago. Despite several decades of research and numerous hypotheses about their functions, exactly how synaptophysin and synaptogyrin participate in the synaptic vesicle cycle remains unknown. The discovery of synaptic vesicles Early evidence supporting the existence of synaptic vesicles came from seminal work by Bernard Katz and colleagues, who were studying synaptic transmission at the frog neuromuscular junction (NMJ). Fatt and Katz (1951, 1952) noticed the existence of small, spontaneous membrane depolarizations on the order of 0.5-­‐1.0 mV that occurred 10 in the absence of nerve stimulation and did not require extracellular calcium. These events were subsequently named spontaneous miniature end-­‐plate potentials (mEPPs), to distinguish them from the large stimulation-­‐evoked responses known as end-­‐plate potentials (EPPs). Interestingly, as the calcium concentration in the saline bath was reduced, the size of an EPP declined until its amplitude reached that of a mEPP (Fatt and Katz, 1952; Del Castillo and Katz, 1954b). At sufficiently low calcium levels, many stimuli completely failed to depolarize the muscle, while those stimuli that succeeded resulted in postsynaptic responses that were integral multiples of the average mEPP amplitude (Fatt and Katz, 1952; Del Castillo and Katz, 1954b). These observations were the foundation of the quantal hypothesis of synaptic transmission, namely that discrete packets (quanta) of neurotransmitter are released with a given probability in response to an action potential, with each quantum contributing about 0.5-­‐1.0 mV to the total EPP amplitude (Del Castillo and Katz, 1954a, b). Ultrastructural evidence in support of the quantal hypothesis arrived shortly thereafter with the discovery of synaptic vesicles through studies utilizing electron microscopy (De Robertis and Bennett, 1955; Palay and Palade, 1955). These structures were abundant, homogeneous, and situated at the presynaptic terminal, making them ideal candidates for the neurotransmitter-­‐containing packets predicted by the quantal hypothesis (Fatt and Katz, 1952). Studies using freeze-­‐fracture techniques in conjunction with electron microscopy were able to capture the extremely rapid release of synaptic vesicles and thereby correlate synaptic vesicle exocytosis with nerve stimulation (Heuser et al., 1974; Heuser et al., 1979). Moreover, the development of purification techniques to isolate presynaptic nerve terminals (synaptosomes) and 11 synaptic vesicles allowed for more in-­‐depth analysis of these organelles and their protein and lipid components (De Robertis et al., 1962; Gray and Whittaker, 1962; Whittaker et al., 1964). Despite these technical advances, it took decades to isolate and clone the first integral synaptic vesicle protein – synaptophysin (Jahn et al., 1985; Wiedenmann and Franke, 1985; Buckley et al., 1987; Leube et al., 1987; Sudhof et al., 1987). Since the discovery of synaptophysin, dozens of proteins associated with synaptic vesicles have been identified and characterized, making the synaptic vesicle one of the best-­‐described intracellular organelles (Jahn and Sudhof, 1994; Sudhof, 2004). The synaptic vesicle cycle Calcium-­‐dependent exocytosis is a highly regulated process, requiring the coordination of a wide array of proteins. Synaptic vesicles must first traffic to specialized release sites known as active zones, where a subset of vesicles dock and are primed for exocytosis (reviewed in (Sudhof, 2004)). During docking and priming, the vesicular v-­‐SNARE synaptobrevin interacts with the plasma membrane t-­‐SNARE syntaxin and SNAP-­‐25 to form the core SNARE complex (Sollner et al., 1993). The association of these proteins results in the formation of a tight, four-­‐helix bundle that brings the vesicle and plasma membrane into close proximity and is thought to assist in overcoming the energetic requirements necessary for vesicle fusion (Hanson et al., 1997; Sutton et al., 1998; Sorensen et al., 2006; Li et al., 2007). However, unlike most other forms of intracellular vesicle trafficking, the formation of the synaptic vesicle SNARE complex does not automatically promote full membrane fusion. Rather, the complex is held in a partial, primed state of fusion until an action potential reaches the 12 presynaptic terminal and calcium entry occurs through voltage-­‐gated calcium channels. Synaptotagmin, as the major calcium sensor in the presynaptic terminal, plays a critical role in coordinating calcium influx with synaptic vesicle fusion (Geppert et al., 1994; Fernandez-­‐Chacon et al., 2001). Furthermore, the protein complexin associates with the intact SNARE complex (McMahon et al., 1995), and while its exact exocytic function is still being elucidated, complexin has been shown to both inhibit and facilitate synaptic vesicle fusion depending on the model organism and assay used (Reim et al., 2001; Giraudo et al., 2006; Huntwork and Littleton, 2007; Xue et al., 2007; Cho et al., 2010). After synaptic vesicles have fused with the presynaptic plasma membrane in response to an action potential, synaptic vesicle proteins and lipids are retrieved through one of three endocytic pathways (Figure 1). First, they may be endocytosed by way of traditional clathrin-­‐mediated endocytosis, often at locations somewhat removed from the sites of exocytosis (Heuser and Reese, 1973; Gad et al., 1998; Slepnev and De Camilli, 2000; Murthy and De Camilli, 2003). Second, synaptic vesicles and their components may be recycled intact through a mechanism called “kiss-­‐and-­‐run” endocytosis, which is proposed to result from the direct fission of a synaptic vesicle immediately following neurotransmitter release and prior to full collapse of the vesicle into the plasma membrane (Ceccarelli et al., 1973; Breckenridge and Almers, 1987; Fesce et al., 1994; Valtorta et al., 2001). Finally, multiple synaptic vesicles may bud from large endocytic cisternae that are formed through a mechanism known as bulk endocytosis (Richards et al., 2000; Evans and Cousin, 2007; Cousin, 2009). Regardless of the endocytic pathway used to regenerate a synaptic vesicle, once it has formed it 13 must be refilled with neurotransmitter, which is accomplished by various transporters that utilize an electrochemical gradient generated by a vacuolar proton pump (Maycox et al., 1988; Sudhof, 2004). After a synaptic vesicle has been refilled with neurotransmitter, it is ready to re-­‐enter one of several synaptic vesicle pools, and this is discussed in more detail below. Clathrin-­mediated vs. kiss-­and-­run endocytosis Efficient synaptic vesicle endocytosis is necessary to maintain a pool of synaptic vesicles that can continually sustain neurotransmission. Early ultrastructural analysis led to two alternative hypotheses as to the nature of synaptic vesicle retrieval. One hypothesis proposed by Heuser and Reese stated that synaptic vesicles fully collapse into the plasma membrane during neurotransmitter release and are later retrieved via clathrin-­‐coated vesicles (1973). Ceccarelli and colleagues suggested a different mechanism in which synaptic vesicles directly reform following release without proceeding through a clathrin-­‐coated intermediate (1973). This method of vesicle exo-­‐
endocytosis was later named “kiss-­‐and-­‐run” and proposes the formation of a narrow fusion pore that allows for transient neurotransmitter release and prevents significant mixing of membrane lipids during exocytosis (Fesce et al., 1994; Stevens and Williams, 2000). Kiss-­‐and-­‐run vesicle retrieval provides certain theoretical advantages over traditional clathrin-­‐mediated endocytosis, namely that vesicles could recycle more quickly and they could retain a full complement of synaptic vesicle proteins and lipids without having to undergo any subsequent sorting steps. Some studies suggest a modified version of kiss-­‐and-­‐run called “kiss-­‐and-­‐stay,” in which vesicles remain apposed to the plasma membrane and are therefore rapidly available for re-­‐use once 14 they have been reloaded with neurotransmitters (Pyle et al., 2000; Stevens and Williams, 2000; Sudhof, 2000). While evidence in support of clathrin-­‐mediated endocytosis at synapses is convincing, the existence and relative contribution of kiss-­‐and-­‐run endocytosis in neurons remains under intense debate (Valtorta et al., 2001; He and Wu, 2007; Rizzoli and Jahn, 2007). This dispute has focused specifically on synaptic vesicles because there is widely accepted evidence in support of kiss-­‐and-­‐run endocytosis in large secretory cells such as mast and chromaffin cells. In these cell types, capacitance measurements and amperometric recordings can detect the rapid opening and closing of fusion pores, which can result in the partial discharge of secretory transmitters (Fernandez et al., 1984; Oberhauser and Fernandez, 1996; Ales et al., 1999; Wang et al., 2003). However, demonstrating the existence of kiss-­‐and-­‐run endocytosis at synapses is more difficult since it relies on indirect methods of observation that can lead to conflicting interpretations (He and Wu, 2007). For example, two groups examining mutations of the endocytic protein endophilin at the Drosophila neuromuscular junction observed the same electrophysiological phenomenon, namely that sustained synaptic transmission was dramatically inhibited (~80-­‐90%) in the endophilin mutant, although a low level of synaptic vesicle release was maintained indefinitely (Verstreken et al., 2002; Dickman et al., 2005). One group (Verstreken et al., 2002) attributed this residual fusion to kiss-­‐and-­‐run endocytosis, while the other (Dickman et al., 2005) concluded that clathrin-­‐mediated endocytosis was crippled but not entirely abolished in the endophilin mutant. Work in hippocampal cells suggests that clathrin-­‐mediated endocytosis is the dominant endocytic pathway under mild stimulation conditions, as 15 the inhibition of this pathway using RNA interference against clathrin or the overexpression of a dominant-­‐negative clathrin accessory protein (AP-­‐180) resulted in a complete loss of endocytosis (Granseth et al., 2006). A more recent study used the FlAsH-­‐FALI technique to photoinactivate clathrin and found that complete inactivation of clathrin-­‐mediated endocytosis resulted in a total loss of synaptic vesicles, again arguing against a clathrin-­‐independent version of kiss-­‐and-­‐run endocytosis (Heerssen et al., 2008). These studies suggest that, even if kiss-­‐and-­‐run endocytosis does occur under certain circumstances at these synapses, clathrin-­‐mediated endocytosis is still the dominant synaptic vesicle retrieval mechanism under most physiological stimulation conditions (although see the discussion on bulk endocytosis below). Clathrin-­mediated endocytosis Clathrin-­‐mediated endocytosis of synaptic vesicles is analogous to other ubiquitous forms of intracellular trafficking that involve clathrin-­‐coated vesicles (Brodin et al., 2000). Heuser and Reese originally proposed that clathrin-­‐mediated endocytosis of synaptic vesicles occurred at distinct sites from exocytosis (1973), and indeed, at certain synapses such as the Drosophila NMJ, proteins associated with clathrin-­‐mediated endocytosis appear to concentrate in specialized “endocytic zones” that surround active zones (Gad et al., 1998; Roos and Kelly, 1999; Dunaevsky and Connor, 2000; Rizzoli and Jahn, 2007). Clathrin-­‐mediated endocytosis is initiated when a clathrin triskelion interacts with adaptor proteins on the plasma membrane to form a protein lattice that covers the invaginating vesicle (Smith and Pearse, 1999). At the synapse, the assembling coat relies on adaptor-­‐binding protein 2 (AP-­‐2) to link clathrin to the membrane (Slepnev and De Camilli, 2000). AP-­‐2 in turn associates with proteins 16 (such as synaptotagmin 1) that contain tyrosine-­‐based targeting sequences and certain acidic phospholipids such as phosphoinositides, thereby linking clathrin coat formation with specific synaptic vesicle proteins and lipids (Gaidarov and Keen, 1999; Haucke and De Camilli, 1999; von Poser et al., 2000). While purified clathrin and AP-­‐2 alone can form cages that resemble clathrin coats, many accessory proteins help mediate clathrin assembly and disassembly and can thereby dramatically affect endocytosis at the synapse (reviewed in (Hirst and Robinson, 1998; Slepnev and De Camilli, 2000)). For example, synaptic vesicle size is regulated at least in part by the neuronal clathrin adaptor protein AP-­‐180, as the loss of this protein in Caenorhabditis elegans or Drosophila causes an increase in synaptic vesicle size and heterogeneity (Zhang et al., 1998a; Nonet et al., 1999). Similarly, in vitro clathrin assembly assays result in smaller, more homogeneous vesicles in the presence of AP-­‐180 (Ye and Lafer, 1995). Misregulation of synaptic vesicle size can directly impact synaptic transmission, as quantal size is larger and more variable in the Drosophila AP-­‐180 mutant (also known as lap) (Zhang et al., 1998a). AP-­‐180 appears to promote clathrin coat assembly through interactions with both clathrin and phosphoinositides (Ahle and Ungewickell, 1986; Hao et al., 1997), and may also play a role in targeting synaptobrevin, as this protein is mislocalized in the C. elegans knockout (Nonet et al., 1999). Another important clathrin accessory protein is synaptojanin, a poly-­‐
phosphoinositide phosphatase that localizes to coated endocytic vesicles (McPherson et al., 1996; Haffner et al., 1997). Synaptojanin plays a critical role in controlling the phosphorylation state of phosphoinositides such as polyphosphatidylinositol-­‐4,5-­‐
17 bisphosphate (PI(4,5)P2), a phospholipid that is thought to promote endocytosis via interactions with other endocytic proteins such as synaptotagmin, AP-­‐2, AP-­‐180, and clathrin (Takei and Haucke, 2001). Like AP-­‐180, synaptojanin appears to directly or indirectly regulate synaptic vesicle size, as synaptic vesicle diameter is increased in the Drosophila mutant (Dickman et al., 2005). The loss of synaptojanin also leads to an accumulation of clathrin-­‐coated vesicles, which suggests that synaptojanin promotes the uncoating of synaptic vesicles following endocytosis (Cremona et al., 1999; Harris et al., 2000; Verstreken et al., 2003). Endophilin, which contains an N-­‐terminal BAR (Bin/Amphiphysin/Rvs-­‐
homology) domain and a C-­‐terminal SH3 domain, appears to play a key role in endocytosis (Ringstad et al., 1997; Schuske et al., 2003). The BAR domain forms a rigid, crescent shape that promotes membrane curvature and induces lipid tubulation in vitro (Farsad et al., 2001; Masuda et al., 2006). Endophilin was also hypothesized to promote membrane curvature by combining lysophosphatidic acid and acyl-­‐CoA to create phosphatidic acid, which has a more conical shape and could induce asymmetry into the lipid bilayer (Schmidt et al., 1999). However, this lysophosphatidic acid acyl transferase activity was later demonstrated to be an artifact of purification (Gallop et al., 2005), so it appears as though endophilin’s membrane-­‐bending activity is merely structural in nature. Endophilin also binds synaptojanin and dynamin through its SH3 domain, and may therefore play a role in recruiting other endocytic proteins (Ringstad et al., 1997). The loss of endophilin results in phenotypes that resemble those of synaptojanin mutants and causes a severe depletion of synaptic vesicles as well as slow endocytic recovery after stimulation (Harris et al., 2000; Verstreken et al., 2002; Dickman et al., 18 2005). Interestingly, overexpression of endophilin can partially rescue a synaptojanin hypomorphic mutant, which lends support to the idea that these proteins function together to promote synaptic vesicle uncoating (Verstreken et al., 2003). The importance of dynamin in synaptic vesicle endocytosis was elegantly revealed in experiments utilizing a temperature-­‐sensitive allele of the Drosophila dynamin homolog named shibire (Koenig and Ikeda, 1989). At the restrictive temperature, endocytosis ceases and shibire mutants gradually lose synaptic vesicles until evoked and spontaneous release arrest and the animals become paralyzed (Grigliatti et al., 1973; Koenig and Ikeda, 1983). Ultrastructural analysis at this point reveals a lack of synaptic vesicles and the existence of numerous membrane invaginations attached to the plasma membrane by a thin neck surrounded by an electron-­‐dense substance (Koenig and Ikeda, 1989). When the endocytic block is released by a shift to the permissive temperature, the synaptic vesicle pool begins to recover, demonstrating that dynamin activity is essential for a late step in endocytosis. Although the precise mechanisms are unclear, dynamin appears to promote synaptic vesicle scission using the energy generated by GTP hydrolysis to alter its conformation and mechanically induce fission (Sweitzer and Hinshaw, 1998; Stowell et al., 1999; Roux et al., 2006; Bashkirov et al., 2008). Bulk endocytosis Bulk endocytosis differs from clathrin-­‐mediated endocytosis in that it does not result in the formation of a single vesicle, but rather generates multiple vesicles through a two-­‐step process (Richards et al., 2000; Cousin, 2009). First, large regions of the plasma membrane are internalized to form endocytic cisternae, which are large 19 intracellular structures that resemble vacuoles or endosomes (Miller and Heuser, 1984; Takei et al., 1996). Next, synaptic vesicles are regenerated from these cisternae, with at least some of these vesicles forming via clathrin-­‐mediated endocytosis (Miller and Heuser, 1984; Takei et al., 1996; Gad et al., 1998; Richards et al., 2000). Bulk endocytosis provides a mechanism through which the endocytic capacity of a presynaptic terminal can be enhanced and is specifically triggered by stimuli that overwhelm the traditional clathrin-­‐mediated pathway, which appears to have a limited capacity (Sankaranarayanan and Ryan, 2000; Clayton et al., 2008; Cousin, 2009). These stimuli can be both non-­‐physiological, such as incubations with solutions containing 4-­‐
aminopyridine or high concentrations of potassium (Miller and Heuser, 1984; Takei et al., 1996; Akbergenova and Bykhovskaia, 2009), or more physiological such as an action potential tetanus (Paillart et al., 2003; Clayton et al., 2008). These forms of intense stimulation can generate endocytic cisternae within a few seconds (Miller and Heuser, 1984; Teng et al., 2007), while clathrin-­‐mediated endocytosis has a time constant of ~14-­‐15 seconds (Granseth et al., 2006). Furthermore, uptake experiments using the fluid-­‐phase marker horseradish peroxidase (HRP) suggest that bulk endocytosis ceases almost immediately following intense stimulation, as very little HRP is internalized into cisternae if it is added after stimulation ends (Clayton et al., 2008). Bulk endocytosis therefore provides an extremely rapid and precise mechanism for reclaiming synaptic vesicle proteins and lipids from the plasma membrane following excessive exocytosis. Compared to clathrin-­‐mediated endocytosis, relatively little is known about the molecular mechanisms behind bulk endocytosis, including how it is initiated. One intriguing candidate for the activity-­‐dependent endocytic trigger is the calcium-­‐
20 dependent phosphatase calcineurin, which may link the elevated calcium levels induced by intense stimulation to the induction of bulk endocytosis (Cousin, 2009). Work in primary neuronal cultures has demonstrated that the inhibition of calcineurin dephosphorylation activity results in the arrest of bulk endocytosis (Evans and Cousin, 2007) and that calcineurin dephosphorylates dynamin under stimulation conditions that trigger bulk endocytosis (Clayton et al., 2009). Furthermore, calcineurin is not required for synaptic vesicle biogenesis during mild activity, suggesting that its function is specific for bulk endocytosis triggered by intense stimulation (Kumashiro et al., 2005; Evans and Cousin, 2007; Clayton and Cousin, 2008). Synaptic vesicle pools Aside from the small minority of vesicles that are morphologically docked at active zones, synaptic vesicles appear relatively indistinguishable by electron microscopy. Nevertheless, synaptic vesicles can be separated into groups based on their functional properties, most notably by their varying propensities to undergo exocytosis (reviewed in (Rizzoli and Betz, 2005)). The definitions and exact names vary, but in general synaptic vesicles can be broken down into three pools: the readily releasable pool (RRP), the recycling pool, and the reserve pool. The RRP comprises those synaptic vesicles that are released immediately upon stimulation with short action potential bursts (Elmqvist and Quastel, 1965; Delgado et al., 2000; Richards et al., 2003). This also is the pool thought to represent the synaptic vesicles docked at release sites (Schikorski and Stevens, 1997, 2001). The recycling pool is defined as the vesicles that are released and reclaimed by endocytosis under normal physiological stimulation conditions (Harata et al., 2001; Kuromi and Kidokoro, 2003; Richards et al., 2003), 21 while the reserve pool is only mobilized by intense or even non-­‐physiological stimuli, such as prolonged high-­‐frequency stimulation or the application of high-­‐potassium solutions (Heuser and Reese, 1973; Delgado et al., 2000; Kuromi and Kidokoro, 2000; Richards et al., 2000; de Lange et al., 2003). In some preparations, including much of the work done at the Drosophila larval NMJ, the RRP and recycling pool are combined and referred to as the exo-­‐endo cycling pool (Kuromi and Kidokoro, 1998, 2002, 2003). In general, the vast majority of synaptic vesicles at a synapse are part of the reserve pool, which can encompass ~80-­‐90% of vesicles, while most of the remaining vesicles are found in the recycling pool (~15-­‐20%), with only a small minority (~1-­‐2%) making up the RRP (Rizzoli and Betz, 2005). At the Drosophila NMJ, analysis of synaptic vesicle release kinetics using the shibire dynamin mutant to block endocytosis similarly found that out of a total of ~84,000 quanta released during synaptic vesicle depletion, about 15-­‐20% were part of the recycling pool and ~300 quanta made up the rapidly depleted RRP (Delgado et al., 2000). The use of styryl dyes such as FM1-­‐43 and FM2-­‐10 has greatly facilitated the study of synaptic vesicle pools and recycling dynamics. These dyes dramatically increase their quantum yield upon binding lipid membranes and therefore can be used to track the movement of vesicles and cisternae through the exo-­‐endocytic cycle (Cochilla et al., 1999; Verstreken et al., 2008). Due to their slightly altered affinities for synaptic lipids, different FM dyes preferentially label distinct endocytic pathways – FM1-­‐43 is taken up by both clathrin-­‐mediated endocytosis and bulk endocytosis, while FM2-­‐10 is largely excluded from endocytic cisternae and therefore mostly labels clathrin-­‐mediated endocytosis (Richards et al., 2000; Evans and Cousin, 2007; Clayton 22 and Cousin, 2008; Clayton et al., 2008). A recent study in neuronal cultures compared the loading of these dyes under stimulation conditions that induced either clathrin-­‐
mediated or bulk endocytosis to investigate how different synaptic vesicle pools were replenished by these two endocytic pathways (Cheung et al., 2010). Intriguingly, clathrin-­‐mediated endocytosis preferentially refilled the recycling pool, while synaptic vesicles generated by bulk endocytosis were targeted to the reserve pool (Figure 1). These findings corroborated previous studies from the frog neuromuscular junction, which found that synaptic vesicles internalized via bulk endocytosis were not immediately available for release, suggesting that they were sent to the reserve pool (Richards et al., 2000; Richards et al., 2004). The SNARE complex and regulated synaptic vesicle fusion Liposome fusion assays have suggested that synaptobrevin, SNAP 25, and syntaxin, the core components of the SNARE complex, are sufficient to bring about membrane fusion (Weber et al., 1998). It should be noted, however, that the in vivo relevance of these experiments is debated, in part because liposome fusion was inefficient, with lipid mixing taking minutes to hours, and the protein/lipid ratios on liposomes were exceptionally high (~750 copies of synaptobrevin per liposome)(Rizo et al., 2006). Further analysis found that the SNARE protein concentration and the method of liposome preparation were critical determinants of lipid-­‐mixing efficiency (Chen et al., 2006). Strikingly, an assay utilizing single molecule fluorescence demonstrated that, while the formation of one SNARE complex is sufficient for liposome docking, fusion was rare even with an average of 12 SNARE complexes per liposome, a value well within the estimated range of complexes necessary for synaptic vesicle 23 fusion (Bowen et al., 2004; Montecucco et al., 2005). These and other observations make it clear that efficient in vivo synaptic vesicle exocytosis requires additional components beyond the SNARE complex. Sec1/Munc18 (SM) proteins participate in a variety of intracellular trafficking events, and many appear to interact specifically with syntaxins (Gallwitz and Jahn, 2003). Mammalian Munc18-­‐1 is essential for neurotransmission, as the loss of this protein results in a complete absence of both spontaneous and evoked synaptic vesicle fusion, even following treatment with α-­‐latrotoxin that induces massive neurotransmitter release in control animals (Verhage et al., 2000). Strikingly, the loss of Munc18 affects exocytosis to an even greater extent than the loss of synaptobrevin, one of the core components of the SNARE complex, indicating that Munc18 plays a critical role in promoting vesicle release (Verhage et al., 2000). Munc18 binds syntaxin in what is known as the “closed” conformation, where an N-­‐terminal region containing a three-­‐helix bundle folds back onto the region of syntaxin that interacts with the other SNARE proteins, thereby preventing SNARE complex formation (Fernandez et al., 1998; Dulubova et al., 1999). It is currently unknown if Munc18 promotes fusion indirectly via its interactions with the SNARE complex or whether it participates in exocytosis directly. Synaptotagmin and complexin Fast, calcium-­‐dependent exocytosis necessarily requires a mechanism to rapidly couple calcium entry through voltage-­‐gated channels to the fusion of synaptic vesicles. Moreover, mechanisms must be in place to ensure that synaptic vesicles do not 24 indiscriminately fuse in the absence of calcium. Synaptotagmin and complexin are two proteins that play critical roles in orchestrating the timing of release and ensuring that vesicle fusion is tied to calcium influx. Synaptotagmin contains two cytoplasmic calcium-­‐binding domains (C2A and C2B) that are homologous to the regulatory regions of protein kinase C and bind three and two calcium ions, respectively (Perin et al., 1990; Ubach et al., 1998; Fernandez et al., 2001). Calcium binding causes synaptotagmin to interact with negatively charged phospholipids, which provide additional coordination sites and thereby increase the apparent calcium affinity of the C2 domains (Zhang et al., 1998b; Fernandez-­‐Chacon et al., 2001). Interestingly, the C2B domain may form a calcium-­‐dependent bridge between synaptic vesicles and the plasma membrane, as it has been shown to simultaneously bind two membranes and bring them into close proximity (Arac et al., 2006). In addition to binding phospholipids, synaptotagmin also interacts with the SNARE complex, although this interaction is only partially dependent on calcium (Bennett et al., 1992; Chapman et al., 1995; Li et al., 1995). As expected, the loss of synaptotagmin severely affects calcium-­‐dependent exocytosis: both Drosophila and mouse synaptotagmin mutants have a dramatic decline in the amplitude of evoked responses, while spontaneous release frequency is enhanced in the Drosophila mutant and the mouse synaptotagmin 2 mutant (but not in the mouse synaptotagmin 1 mutant) (Littleton et al., 1993; Geppert et al., 1994; Pang et al., 2006). Complexins are small, cytosolic proteins containing an α-­‐helical region that tightly binds a groove in the four-­‐helix bundle of the assembled SNARE complex (McMahon et al., 1995; Bracher et al., 2002; Chen et al., 2002). This finding, along with the observation that complexin can inhibit liposome or cell-­‐cell fusion, led to the 25 hypothesis that complexin acts as a fusion clamp by holding SNARE complexes in a hemifused state (Giraudo et al., 2006; Schaub et al., 2006). Additional evidence in support of this hypothesis came from analysis of a Drosophila complexin mutant, which displayed a dramatic increase in the frequency of spontaneous synaptic vesicle release (Huntwork and Littleton, 2007). However, conflicting results were observed in mice, as synapses in the complexin knockout mouse showed a decrease in spontaneous fusion (Reim et al., 2001; Xue et al., 2007; Xue et al., 2008), while knockdown using RNA interference caused an increase in spontaneous release (Maximov et al., 2009). Evoked neurotransmitter release is inhibited in both Drosophila and mouse knockouts, suggesting that complexin can also facilitate release (Reim et al., 2001; Huntwork and Littleton, 2007; Xue et al., 2008). The discrepancies in the data imply that complexin function varies between species, and indeed there appear to be differences between the specific isoforms (Cho et al., 2010). Cross-­‐species rescue and overexpression experiments confirm that Drosophila complexin has a mostly inhibitory role, while murine complexin has a primarily facilitating effect (Xue et al., 2009). Furthermore, in-­‐
depth structure-­‐function analysis has shown that different regions within a single complexin isoform can either promote or inhibit vesicle fusion (Xue et al., 2007; Xue et al., 2009). Therefore, it remains to be determined exactly how complexin regulates synaptic vesicle exocytosis through its interactions with the SNARE complex, as well as how synaptotagmin and complexin might work together to coordinate calcium-­‐
dependent fusion. 26 Molecular anatomy of synaptic vesicles Much of our understanding of the synaptic vesicle cycle has come from the identification and characterization of the proteins located on synaptic vesicles. Because synaptic vesicles are homogeneous in size and density, they can readily be purified from brain homogenates. Recent proteomic studies have utilized a variety of gel electrophoresis techniques combined with mass spectrometry in an attempt to ascertain the full complement of synaptic vesicle proteins (Blondeau et al., 2004; Burre et al., 2006; Takamori et al., 2006). While their findings differ in terms of the exact number and identity of synaptic vesicle proteins, all of these analyses conclude that several hundred different proteins either reside on or associate with synaptic vesicles. One of these studies (Takamori et al., 2006) calculated the protein stoichiometry for synaptic vesicles using quantitative Western blotting and found that the two most abundant proteins were synaptobrevin (~70 copies/vesicle) and synaptophysin (~32 copies/vesicle), while synaptogyrin had approximately two copies per vesicle. In terms of mass, however, synaptophysin is the most abundant, accounting for ~10% of total synaptic vesicle protein (Takamori et al., 2006). Analysis of single isolated synaptic vesicles using total internal reflection fluorescence (TIRF) microscopy gave different copy numbers per vesicle, most notably for synaptobrevin, with averages of about 10 synaptobrevins, 13 synaptophysins, and 7 synaptogyrins per vesicle (Mutch et al., 2011). Although the exact values vary between these two studies, it is clear that both synaptophysin and synaptobrevin are among the most abundant synaptic vesicle proteins. 27 Initial characterization of synaptophysin and synaptogyrin Synaptophysin Synaptophysin, initially named p38 due to its apparent molecular mass of 38 kDa, was among the first synaptic vesicle proteins to be discovered and characterized in detail (Jahn et al., 1985; Wiedenmann and Franke, 1985). The name synaptophysin comes from the combination of “synapse” and the Greek “φῦσα” (bubble or vesicle), and was given due to the protein’s location on synaptic vesicles at the presynaptic terminal (Wiedenmann and Franke, 1985). Due to its ubiquitous presence on synaptic vesicles, synaptophysin has historically been used as a marker to measure synaptic density or diagnose neoplasms, among other uses (Valtorta et al., 2004). Synaptophysin is an N-­‐
glycosylated tetraspanning membrane protein with cytoplasmically located N-­‐ and C-­‐
termini (Figure 2) (Jahn et al., 1985; Wiedenmann and Franke, 1985; Leube et al., 1987; Sudhof et al., 1987). Synaptophysin can form noncovalent homomultimeric complexes ranging in size from dimers to hexamers, and its oligomerization state is not dependent on its C-­‐terminus (Wiedenmann and Franke, 1985; Rehm et al., 1986; Thomas et al., 1988; Johnston and Sudhof, 1990; Pennuto et al., 2002; Arthur and Stowell, 2007). Each of synaptophysin’s two luminal domains contains a disulfide bond that forms within each intravesicular loop and does not interact with neighboring synaptophysin subunits in a complex (Johnston and Sudhof, 1990). Synaptophysin’s C-­‐terminus has numerous glycine and proline residues and contains ten pentapeptide repeats, nine of which have a tyrosine residue (Sudhof et al., 1987). The large number of tyrosine residues suggests that synaptophysin can undergo phosphorylation, and indeed synaptophysin is a major target of pp60c-­‐src tyrosine kinase, which is also associated with synaptic vesicles (Pang 28 et al., 1988; Barnekow et al., 1990; Linstedt et al., 1992). The interaction between synaptophysin and pp60c-­‐src is enhanced following training with the Morris water maze spatial learning task, suggesting that synaptophysin function may be modulated during learning and memory (Zhao et al., 2000). Similarly, tyrosine phosphorylation of synaptophysin was increased in hippocampal brain slices that had undergone long-­‐
term potentiation, suggesting that synaptophysin may play a role in synaptic plasticity (Mullany and Lynch, 1998). Synaptophysin also appears to be regulated via serine phosphorylation, as it has been shown to be a calcium-­‐dependent substrate of CaM kinase II (Rubenstein et al., 1993). This raises the possibility that synaptophysin function may be modulated in response to the calcium influx triggered by action potentials. While there is strong evidence to suggest that synaptophysin is phosphorylated under physiological conditions, the locations of the tyrosine and serine residues that undergo phosphorylation are currently unknown (Evans and Cousin, 2005). Synaptogyrin Synaptogyrin was originally identified as a 29 kDa protein located throughout the nervous system as well as endocrine cells that have regulated secretion (Baumert et al., 1990). Like synaptophysin, synaptogyrin received its name due to its presence on synaptic vesicles (the Greek “γῦρος” means circle)(Stenius et al., 1995). Synaptogyrin also contains four transmembrane domains and has the same membrane orientation as synaptophysin (Stenius et al., 1995). Furthermore, like synaptophysin, synaptogyrin has several tyrosine residues in its C-­‐terminus that are phosphorylated by pp60c-­‐src (Figure 2) (Janz and Sudhof, 1998). When transfected into COS cells, synaptogyrin and 29 synaptophysin are both phosphorylated by pp60c-­‐src and c-­‐fyn, with synaptogyrin appearing to be the better target for phosphorylation by c-­‐fyn (Janz et al., 1999). However, unlike synaptophysin, synaptogyrin is not glycosylated, and only its first intravesicular domain has a pair of conserved cysteine residues that potentially form a disulfide bond (Baumert et al., 1990; Stenius et al., 1995). The C-­‐terminus of synaptogyrin is required for proper targeting of the protein to synapses in C. elegans and is also necessary for the correct localization of GFP-­‐tagged rat synaptogyrin in cultured hippocampal neurons (Zhao and Nonet, 2001). Unlike synaptophysin, the quaternary structure of synaptogyrin has not been investigated (Hubner et al., 2002). Since relatively few studies have examined synaptogyrin, much less is known about the structure and function of synaptogyrin compared to synaptophysin. Physins and gyrins: evolutionarily related MARVEL domain proteins Once the amino acid sequences of synaptophysin and synaptogyrin were determined, it became apparent that these proteins were distantly related through evolution (Sudhof et al., 1987; Stenius et al., 1995). Moreover, each of these proteins became the founding member of two larger protein families: the physins and the gyrins (Hubner et al., 2002). In mammals, the physin family encompasses synaptophysin, synaptoporin (also called synaptophysin 2), pantophysin, and mitsugumin 29, while the synaptogyrin family consists of synaptogyrins 1-­‐4 (synaptogyrin 2 is also referred to as cellugyrin). All of these proteins contain a conserved tetraspanning membrane motif called the MARVEL domain, for MAL and related proteins for vesicle trafficking and membrane link (Sanchez-­‐Pulido et al., 2002). However, the composition and length of 30 the N-­‐ and C-­‐termini of physins and gyrins varies significantly, even within a family (Hubner et al., 2002). As the name suggests, MARVEL domain-­‐containing proteins have been found to regulate membrane associations and intracellular trafficking. For example, the myelin and lymphocyte protein (MAL) and the related protein MAL2 mediate apical transport in polarized cells. MAL regulates the transport of cargo from the trans-­‐Golgi network to the apical membrane in epithelial cells (Cheong et al., 1999; Puertollano and Alonso, 1999), while MAL2 participates in transcytosis in hepatocytes (de Marco et al., 2002). Both MAL and MAL2 are present in lipid rafts, raising the possibility that these proteins mediate apical transport by recruiting appropriate lipids and proteins through their transmembrane domains (Cheong et al., 1999; Puertollano and Alonso, 1999; de Marco et al., 2002). Interestingly, MAL (but not MAL2) has been shown to oligomerize and recruit apical proteins in a cholesterol-­‐dependent manner (Magal et al., 2009; Ramnarayanan and Tuma, 2011). MAL’s ability to promote and stabilize lipid rafts may lie in its relatively long transmembrane helices (23-­‐25 amino acids), which may specifically recruit proteins and lipids destined for the apical membrane, since the apical surface is thicker than other cellular membranes (Magal et al., 2009; Ramnarayanan and Tuma, 2011). Plasmolipin and BENE, two other members of the MAL protein family, are also associated with lipid rafts (de Marco et al., 2001; Bosse et al., 2003). Plasmolipin is enriched in myelin and is linked to the pleiotropic disorder Bardet-­‐Biedl syndrome (Hamacher et al., 2001), while BENE has been implicated in intracellular trafficking in endothelial cells (de Marco et al., 2001). BENE also interacts with caveolin, a protein known to organize lipid rafts during the formation of caveolae – 31 the plasma membrane invaginations involved in diverse processes such as cell signaling and lipid regulation (Parton and Simons, 2007). A second family of MARVEL proteins comprising occludin, tricellulin, and the newly characterized protein marvelD3 are targeted to tight junctions, which are the structures that separate apical and basolateral membranes and form a barrier to prevent the movement of solutes between epithelial cell layers (Furuse et al., 1993; Tsukita et al., 2001; Raleigh et al., 2010). However, occludin is not required for the formation of functional tight junctions, since occludin mice, while they are slow to develop postnatally and have several histological phenotypes, are nevertheless able to form effective epithelial barriers (Saitou et al., 2000). Although the exact role of the occludin family is unknown, all three proteins are present in detergent-­‐insoluble membrane microdomains that are enriched in cholesterol and sphingolipids, suggesting that they may play a role in organizing the lipid rafts around tight junctions (Nusrat et al., 2000; Raleigh et al., 2010). Interestingly, overexpression of chicken occludin in insect cells (which normally do not form tight junctions) causes the formation of intracellular multilamellar structures with connections resembling tight junctions, suggesting that occludin has the ability to induce membrane reorganization (Furuse et al., 1996; Furuse and Tsukita, 2006). In Drosophila, the MARVEL protein Singles Bar is involved in regulating myoblast fusion during the formation of multinucleated muscles (Estrada et al., 2007; Abmayr et al., 2008). In the absence of Singles Bar, myoblasts adhere to one another but do not fuse, indicating that the protein is involved in a late step in the fusion process (Estrada et al., 2007). While the exact mechanism is unknown, Singles Bar is proposed to 32 mediate the fusion of dense-­‐core vesicles with the plasma membrane to deliver components required to promote the final steps of myoblast fusion. No vertebrate homologs of Singles Bar have been identified, and currently no other MARVEL domain-­‐
containing homolog has been implicated in myoblast fusion (Estrada et al., 2007). However, caveolin-­‐3 is required for myoblast fusion in mammals (Galbiati et al., 1999), raising the possibility that proper organization of membrane lipids is an essential step for promoting membrane fusion in both systems. Taken together, these findings suggest that the MARVEL domain plays a general role in organizing membrane microdomains and controlling membrane apposition events. The ability of many MARVEL domain-­‐containing proteins to manipulate the lipid composition and architecture of membranes may also underlie the function of gyrins and physins. Synaptoporin, pantophysin, and mitsugumin 29 Synaptophysin and synaptoporin are more closely related to each other than to the other physins pantophysin and mitsugumin 29 (Hubner et al., 2002). Like synaptophysin, synaptoporin is enriched in the brain and neuroendocrine cells (Knaus et al., 1990), although its pattern of expression is more narrow than that of synaptophysin (Hubner et al., 2002). Pantophysin was initially believed to be ubiquitously expressed (Leube, 1994); however, although all major tissues contain pantophysin mRNA, its protein expression is heterogeneous in various cell types (Haass et al., 1996; Windoffer et al., 1999). On the other hand, mitsugumin 29 expression is limited to skeletal muscle, small intestine, and kidney (Shimuta et al., 1998; Komazaki et al., 1999). In skeletal muscle, mitsugumin 29 is enriched at the triad junction, a location 33 where changes in membrane potential are coupled to calcium release from the sarcoplasmic reticulum (Franzini-­‐Armstrong and Jorgensen, 1994; Nishi et al., 1999). Knockout experiments in mice revealed that the loss of mitsugumin 29 causes morphological abnormalities in the membranes surrounding the triad junction, as well as increased fatigue and delayed recovery of certain muscle types, which may stem from altered intracellular calcium homeostasis (Nishi et al., 1999; Nagaraj et al., 2000; Komazaki et al., 2001). Synaptogyrins 2-­4 Cellugyrin, also known as synaptogyrin 2, is ubiquitously expressed, although its mRNA levels are lower in the brain relative to other tissues (Janz and Sudhof, 1998; Kedra et al., 1998). Synaptogyrin 3 mRNA is restricted to the brain and placenta (Kedra et al., 1998); however, synaptogyrin 3 expression in the brain is not as widespread as synaptogyrin 1 (Belizaire et al., 2004). Thus, both synaptophysin and synaptogyrin have a paralog with a more restricted neuronal expression pattern, suggesting that the functions of synaptoporin and synaptogyrin 3 may be more specialized. Synaptogyrin 4 expression has not yet been established (Hubner et al., 2002). SCAMPs Secretory carrier-­‐associated membrane proteins (SCAMPs), like physins and gyrins, are tetraspanning membrane proteins with cytoplasmic N-­‐ and C-­‐termini (Hubbard et al., 2000). While their membrane topology is similar, SCAMPs lack a MARVEL domain and share little sequence homology with physins or gyrins (Hubner et al., 2002; Sanchez-­‐Pulido et al., 2002). All SCAMPs share a conserved domain that 34 encompasses the four transmembrane helices, and each has a highly conserved region located in the cytoplasmic loop between the second and third transmembrane domains called the E peptide (Hubbard et al., 2000). SCAMPs were originally identified as components of secretory vesicles in exocrine glands (Cameron et al., 1986; Brand et al., 1991; Laurie et al., 1993). However, most SCAMPs have since been shown to have a much broader expression pattern and are not restricted to synaptic vesicles (Singleton et al., 1997; Fernandez-­‐Chacon and Sudhof, 2000; Hubner et al., 2002). Unlike gyrins and physins, SCAMP homologs are also found in plants, with the Arabidopsis genome encoding four SCAMP homologs, all of which are closely related to one another (Fernandez-­‐Chacon and Sudhof, 2000). Mammals have five SCAMP isoforms, three of which (SCAMPs 1-­‐3) have multiple NPF repeats in their N-­‐termini, while SCAMPs 4 and 5 lack this domain (Fernandez-­‐Chacon and Sudhof, 2000). Interestingly, SCAMP 5 was only found in the brain where it is highly enriched on synaptic vesicles, suggesting that its function is specific to neurons, unlike the other more ubiquitous SCAMPs that likely have a more general cellular function (Fernandez-­‐Chacon and Sudhof, 2000). The NPF domain has been identified as a binding site for proteins containing an Eps15 homology (EH) domain (de Beer et al., 1998; Paoluzi et al., 1998), a domain that has been implicated in clathrin-­‐mediated endocytosis and vesicle budding from the trans-­‐Golgi complex (Tebar et al., 1996; Chi et al., 2008). SCAMPs also are thought to function in endocytosis in multiple cell types by recruiting EH domain-­‐containing proteins such as intersectin (Fernandez-­‐Chacon et al., 2000), a protein that also has been implicated in endocytosis (Hussain et al., 1999). 35 SCAMPs also have been proposed to regulate exocytosis, as overexpression of the E peptide in both mast cells and PC12 cells inhibits the release of granules and dense-­‐core vesicles, respectively (Guo et al., 2002; Liu et al., 2002). SCAMP1-­‐deficient mice are viable and fertile and do not appear to have major deficits in exocytosis or endocytosis (Fernandez-­‐Chacon et al., 1999). However, capacitance recordings from mast cells revealed an increase in the frequency of transient fusion events, suggesting that SCAMP1 may function to stabilize a fusion pore and/or regulate endocytosis following fusion (Fernandez-­‐Chacon et al., 1999). Interestingly, the SCAMP cytoplasmic loop (the E peptide) and the two cytoplasmic segments bordering transmembrane domains 1 and 4 are amphiphilic, raising the possibility that these regions might interact with lipid bilayers to affect fusion pore dynamics (Hubbard et al., 2000). Potential functions of synaptophysin and synaptogyrin Regulation of exocytosis The observation that synaptophysin forms a multimeric complex has lead to speculation that it might form an ion channel or proteinaceous fusion pore that could allow for rapid neurotransmitter release during exocytosis (Sudhof et al., 1987; Valtorta et al., 2004). Ultrastructural analysis of purified synaptophysin revealed rosette-­‐like particles with a mean diameter of approximately 8 nm (Thomas et al., 1988), while single-­‐particle three-­‐dimensional reconstruction identified a hexameric structure with an outer diameter of about 7 nm and an inner diameter of approximately 3 nm (Arthur and Stowell, 2007). When synaptophysin complexes were reconstituted into planar lipid bilayers, they displayed voltage-­‐sensitive channel activity with an average conductance of 150 picosiemens (Thomas et al., 1988), a value similar to that of gap 36 junctions (Young et al., 1987). These reconstituted channels appear to be highly selective for potassium ions (Gincel and Shoshan-­‐Barmatz, 2002), although whether synaptophysin has the ability to conduct ions in vivo remains to be determined. The hypothesis that synaptophysin creates a fusion pore or an ion channel is enhanced by the observation that synaptophysin and the gap junction protein connexin share a similar membrane topology and overall structure (Leube, 1995; Arthur and Stowell, 2007). Connexin’s third transmembrane domain is predicted to line the pore of gap junctions (Skerrett et al., 2002), and while connexin and synaptophysin have little global sequence homology, their third transmembrane domains are much more similar, raising the possibility that synaptophysin’s third transmembrane domain may, by analogy, line the fusion pore (Arthur and Stowell, 2007). Furthermore, experiments utilizing fluorescence resonance energy transfer (FRET) of CFP-­‐ and YFP-­‐tagged synaptophysin molecules found that synaptophysin oligomers dissociate upon synaptic vesicle exocytosis, which may represent the disassembly of a fusion pore upon synaptic vesicle integration into the plasma membrane (Pennuto et al., 2002). While the idea of synaptophysin forming a fusion pore analogous to gap junctions is intriguing, there are several lines of evidence that argue against this possibility. First, there is debate as to which transmembrane domain of connexin lines the pore of gap junctions. Several groups favor the third transmembrane domain as the major pore-­‐lining helix (Skerrett et al., 2002; Fleishman et al., 2004), while other models predict the first transmembrane domain (Zhou et al., 1997; Kronengold et al., 2003; Maeda et al., 2009). Since there is little sequence homology between connexin and synaptophysin outside of the third transmembrane domain, the similarities 37 between the two proteins are lessened if the first transmembrane helix of connexin lines the pore. Second, gap junctions are formed by the association of two connexin hemichannels located on the plasma membranes of adjacent cells, while a putative synaptophysin fusion pore would connect synaptic vesicles with the plasma membrane. Since both connexin and synaptophysin have cytoplasmic N-­‐ and C-­‐termini, the interaction between connexons is mediated by their extracellular loops, while a presumed association between synaptophysin oligomers would occur through their cytoplasmic domains (Sudhof and Jahn, 1991). In other words, the orientation of the hemichannels would be reversed. Finally, if synaptophysin interacts with another binding partner located on the plasma membrane to form a fusion pore, such a protein has not yet been definitively identified (Valtorta et al., 2004). A synaptophysin binding partner named physophilin was initially a candidate for the plasma membrane component of the fusion pore (Thomas and Betz, 1990). However, physophilin was later shown to be a subunit (Vo) of the vacuolar H+ ATPase, a complex enriched on synaptic vesicles that seems unlikely to be the major plasma membrane component of the fusion pore (Siebert et al., 1994; Galli et al., 1996; Carrion-­‐Vazquez et al., 1998). In spite of the evidence against synaptophysin forming a proteinaceous fusion pore analogous to gap junctions, it remains possible that synaptophysin may promote fusion through alternative mechanisms. Indeed, additional support implicating synaptophysin in regulating calcium-­‐dependent synaptic vesicle exocytosis has come from overexpression studies using Xenopus oocytes. Upon injection with total rat cerebellar mRNA, Xenopus oocytes become capable of secreting glutamate in a calcium-­‐
dependent manner (Alder et al., 1992b). When either synaptophysin antisense 38 oligonucleotides or synaptophysin antibodies are co-­‐injected, this glutamate release is substantially inhibited (Alder et al., 1992b). Furthermore, overexpression of synaptophysin increases both spontaneous and evoked neurotransmitter responses in cultured Xenopus motor neurons (Alder et al., 1995), while the injection of synaptophysin antibodies inhibits release (Alder et al., 1992a). Taken together, these results suggest that synaptophysin promotes calcium-­‐dependent neurotransmitter fusion. However, another group reported that overexpression of synaptophysin or synaptogyrin inhibited calcium-­‐dependent release in neuroendocrine PC12 cells engineered to secrete human growth hormone (Sugita et al., 1999). This same study found that synaptogyrins 1-­‐3 all inhibited secretion to a greater extent than synaptophysin and determined that the C-­‐terminus of gyrins was largely disposable for this inhibition. It should be noted, however, that human growth hormone in PC12 cells is stored in and released from large dense-­‐core vesicles rather than traditional synaptic vesicles (Lowe et al., 1988). While secretion via dense-­‐core vesicles shares many similarities with synaptic vesicle exocytosis (e.g., it is triggered by calcium and inhibited by tetanus toxin), it is unclear whether synaptophysin and synaptogyrin normally localize to dense-­‐core vesicles, raising uncertainty about whether the effects of their overexpression in PC12 cells is physiologically relevant (Wiedenmann and Franke, 1985; Navone et al., 1986; Baumert et al., 1990; Sugita et al., 1999). How might synaptophysin and/or synaptogyrin regulate synaptic vesicle exocytosis? One appealing possibility comes from the observation that synaptophysin interacts with synaptobrevin, a critical component of the SNARE complex and the only 39 SNARE component found on synaptic vesicles (Calakos and Scheller, 1994; Edelmann et al., 1995; Washbourne et al., 1995). The interaction between synaptophysin and synaptobrevin is lost when synaptophysin’s disulfide bonds are reduced, which causes synaptophysin to lose its oligomeric state (Johnston and Sudhof, 1990). Since synaptophysin complexes dissociate upon exocytosis (Pennuto et al., 2002), this implies that the synaptophysin/synaptobrevin interaction is also lost following synaptic vesicle fusion. Furthermore, the interaction between synaptobrevin and synaptophysin precludes the association of synaptobrevin with SNAP-­‐25 and syntaxin (Edelmann et al., 1995), which raises the possibility that synaptophysin may regulate the availability of synaptobrevin in a manner similar to the syntaxin/Munc18 interaction (Hata et al., 1993). The availability of syntaxin to interact with other SNARE components is controlled by Munc18, which appears to mediate a conformational switch in syntaxin that allows it to interact with the other SNAREs (Dulubova et al., 1999). Similarly, synaptophysin may regulate SNARE complex assembly and fusion by controlling the availability of synaptobrevin. Overexpression and knockout studies in cultured neurons suggest that synaptophysin may play a role in targeting synaptobrevin to synaptic vesicles. When synaptobrevin is overexpressed only in the presence of endogenous synaptophysin, a portion of synaptobrevin is mislocalized away from synapses (Pennuto et al., 2003; Gordon et al., 2011). However, co-­‐transfection of exogenous synaptophysin along with synaptobrevin restored the correct localization of synaptobrevin. Similar results were obtained in synaptophysin knockout cultures, with both endogenous and exogenous synaptobrevin showing an increased presence on the plasma membrane in the absence 40 of synaptophysin (Gordon et al., 2011). The synaptobrevin mislocalization appears to be due to a failure to retrieve the protein from the plasma membrane via endocytosis, as the levels of synaptobrevin on the plasma membrane after robust stimulation remain elevated in the synaptophysin mutant cultures compared to controls. The promotion of synaptic vesicle protein sorting by synaptophysin appears to be specific for synaptobrevin, because co-­‐transfection of synaptotagmin and synaptophysin cannot rescue synaptotagmin mislocalization (Pennuto et al., 2003). Regulation of endocytosis Additional lines of evidence suggest that gyrins and physins may play a role in endocytosis and/or vesicle biogenesis. Synaptophysin has been shown to bind cholesterol, and the depletion of cholesterol in PC12 cells results in a preferential decrease in the formation of synaptic-­‐like microvesicles (SLMVs) but not an overall decline in endocytosis (Thiele et al., 2000). Cholesterol is enriched in synaptic vesicles (Breckenridge et al., 1973; Takamori et al., 2006), and it has been proposed that synaptophysin’s multimeric oligomerization may recruit cholesterol to promote the formation of highly curved membranes (Thiele et al., 2000). Cellugyrin has also been shown to induce the formation of SLMVs in PC12 cells (Belfort et al., 2005) and to promote the targeting of synaptophysin to SLMVs (Belfort and Kandror, 2003). Interestingly, the hydrophilic cytoplasmic regions of cellugyrin do not appear to be important for vesicle biogenesis. Instead, certain hydrophobic stretches near the transmembrane domains significantly influence cellugyrin’s ability to promote the formation of small vesicles, potentially by directly influencing membrane curvature (Belfort et al., 2005). Transfection of synaptogyrin also results in increased targeting of 41 synaptophysin to small vesicles (Belfort and Kandror, 2003), although the enhanced synaptophysin targeting appears to be limited to PC12 cells (Belfort et al., 2005). Further evidence implicating synaptophysin in vesicle formation comes from the observation that the C-­‐terminus of synaptophysin interacts with dynamin in a calcium-­‐
dependent manner, with half-­‐maximal binding occurring around 150 µM (Daly and Ziff, 2002). Dynamin plays a critical role in endocytosis by utilizing GTPase activity to constrict the neck of a budding vesicle and induce membrane scission (Hinshaw, 2000). Blocking the interaction of dynamin and synaptophysin by injecting the C-­‐terminus of synaptophysin fused to glutathione S-­‐transferase (GST) results in a depletion of the overall synaptic vesicle pool in the squid giant synapse (Daly et al., 2000). At the same time, the number of clathrin-­‐coated vesicles increases, suggesting that clathrin-­‐
mediated endocytosis is not blocked (Daly et al., 2000). Interestingly, the interaction between synaptophysin and dynamin occurs at calcium concentrations seen in microdomains near active zones (~200-­‐300 µM), where voltage-­‐gated calcium channels are enriched (Pumplin et al., 1981; Llinas et al., 1992; Heidelberger et al., 1994; Daly and Ziff, 2002). Taken together, these results suggest that the synaptophysin-­‐dynamin interaction promotes a clathrin-­‐independent form of endocytosis (potentially a version of kiss-­‐and-­‐run endocytosis) that specifically occurs near sites of exocytosis where calcium concentrations are elevated. These findings also imply that the inhibition of this version of endocytosis results in a compensatory enhancement of clathrin-­‐
mediated endocytosis (Daly et al., 2000; Daly and Ziff, 2002). A recent study provided additional support for a role of synaptophysin in synaptic vesicle recycling through the examination of endocytic kinetics in dissociated 42 hippocampal neurons from synaptophysin mutant mice (Kwon and Chapman, 2011). During sustained stimulation, the knockout cultures were slower to internalize synaptic vesicle proteins tagged with a pH-­‐sensitive GFP reporter. However, endocytosis was unaltered during mild stimulation, suggesting that synaptophysin is only required for efficient endocytosis during intense stimulation. Interestingly, synaptophysin promotes endocytosis both during and after stimulation, but the C-­‐terminus of synaptophysin is required only for synaptic vesicle retrieval during the stimulus and not for endocytosis that occurs once the stimulus ceases (Kwon and Chapman, 2011). These findings suggest that distinct synaptophysin domains regulate different endocytic processes. Synaptophysin and synaptogyrin knockouts Although synaptophysin and synaptogyrin are evolutionarily conserved and abundant synaptic vesicle proteins, the loss of these proteins results in relatively mild phenotypes. Both synaptophysin and synaptogyrin knockout mice are viable and fertile, as is the double mutant (Eshkind and Leube, 1995; McMahon et al., 1996; Janz et al., 1999). Mice lacking synaptophysin initially were reported to have slightly lower levels of synaptobrevin (McMahon et al., 1996). However, subsequent analysis of the synaptophysin/synaptogyrin double knockout revealed no significant changes in synaptobrevin expression or in the expression levels a wide variety of other synaptic vesicle proteins (Janz et al., 1999). Intriguingly, synaptophysin mutants exhibit a decrease in synaptic vesicle density and an increase in the number of clathrin-­‐coated vesicles specifically in photoreceptor cells in the outer plexiform layer of the retina – a region that lacks synaptoporin (McMahon et al., 1996; Spiwoks-­‐Becker et al., 2001). Moreover, the rod photoreceptors in knockout animals have altered synaptic vesicle 43 morphology, with many vesicles misshapen or unusually flat (Spiwoks-­‐Becker et al., 2001). On the other hand, synaptic vesicle morphology appears normal in a cerebellar region that also contains little to no synaptoporin (Fykse et al., 1993; Eshkind and Leube, 1995). Many of the phenotypes observed in the retinas of synaptophysin mutant animals were exacerbated during periods of high activity (i.e., when animals were not exposed to light), indicating that synaptophysin’s function may be specific to certain high-­‐intensity stimulation conditions. Electrophysiological analysis of synaptophysin-­‐deficient mice found no significant changes in excitatory postsynaptic current (EPSC) amplitude, mEPSC frequency, or release probability. However, there was a slight (less than 10%) increase in quantal size (McMahon et al., 1996). Although synaptic vesicle morphology was altered in the retinal cells of the knockout mice, electroretinography revealed no significant differences compared to controls, suggesting that the defects seen by electron microscopy do not severely impair vision (Spiwoks-­‐Becker et al., 2001; Schmitt et al., 2009). Further analysis revealed that synaptophysin mutant mice have no significant changes in paired-­‐pulse facilitation (PPF), post-­‐tetanic potentiation (PTP), or long-­‐term potentiation (LTP)(Janz et al., 1999). Moreover, synaptogyrin-­‐deficient mice have only a slight decrease in PTP, with no changes in other forms of synaptic plasticity. Interestingly, mice lacking both synaptophysin and synaptogyrin display defects in all of these types of short-­‐ and long-­‐term synaptic plasticity, as well as a delay in the recovery from synaptic depression, indicating that there is functional redundancy between these two proteins (Janz et al., 1999). Anatomically, double mutant mice have normal brain architecture, suggesting that these proteins are not required to establish 44 gross neuronal patterning during development (Janz et al., 1999). Behaviorally, synaptophysin-­‐deficient mice display normal levels of locomotor activity; however, they were more exploratory in an enriched open field and displayed reduced novel object recognition (Schmitt et al., 2009). The knockout mice also exhibited reduced performance in spatial learning and memory as assayed by the Morris water maze hidden platform test (Schmitt et al., 2009). Interestingly, synaptophysin has also been implicated in human cognitive ability, as large-­‐scale chromosomal sequencing has identified several variants at the synaptophysin locus in subjects with X-­‐linked mental retardation (Tarpey et al., 2009). However, to date no detailed behavioral assays have been performed on synaptogyrin knockout mice or synaptophysin/synaptogyrin double knockouts. As has been previously discussed, analysis of neuronal physin and gyrin function in mammals is complicated by the presence of multiple isoforms of both proteins in the brain. C. elegans, on the other hand, has one synaptophysin homolog and one synaptogyrin homolog, thereby allowing issues of redundancy between physins and gyrins to be more easily ascertained. Somewhat surprisingly, the initial analysis of a C. elegans synaptophysin/synaptogyrin/SCAMP triple knockout also revealed no significant changes in neuronal morphology, synaptic transmission, or behavior aside from a slight increase in the number of clathrin-­‐coated synaptic vesicles (Spiwoks-­‐
Becker et al., 2001; Abraham et al., 2006). Closer examination of the C. elegans synaptogyrin single mutant identified slight changes in sensitivity to certain drugs affecting GABAergic and cholinergic neurons, subtle alterations in motility, and a slight decrease in synaptobrevin targeting to synapses (Abraham et al., 2011). Interestingly, 45 synaptophysin expression in C. elegans appears to be largely restricted to muscle cells in the pharynx and anal sphincter (Abraham et al., 2006). However, synaptogyrin is expressed in almost all neurons and colocalizes with other synaptic vesicle proteins (Nonet, 1999; Zhao and Nonet, 2001; Abraham et al., 2011), indicating that synaptogyrin is likely to be the predominant synaptic vesicle MARVEL protein in nematodes. Interestingly, an increase in clathrin-­‐coated vesicles is observed in both the C. elegans triple knockout and in mouse retinal photoreceptors that lack both synaptophysin and synaptoporin (Spiwoks-­‐Becker et al., 2001; Abraham et al., 2006). As previously mentioned, inhibition of the dynamin-­‐synaptophysin interaction also resulted in an increase in clathrin-­‐coated vesicles, which was presumably compensating for the disruption of a clathrin-­‐independent endocytic pathway mediated by synaptophysin (Daly et al., 2000; Daly and Ziff, 2002). Several of the defects observed in knockout mice, namely the delay in recovery from synaptic depression in the synaptophysin/synaptogyrin double knockout and the synaptic vesicle morphological abnormalities in retinal cells lacking synaptophysin, point to deficiencies in synaptic vesicle endocytosis (Janz et al., 1999; Spiwoks-­‐Becker et al., 2001). Intriguingly, in C. elegans, the loss of synaptogyrin enhanced the phenotypes observed in synaptojanin, endophilin, and synaptotagmin mutants, all of which are regulators of clathrin-­‐mediated endocytosis (Song and Zinsmaier, 2003; Shupliakov, 2009; Abraham et al., 2011). If synaptophysin and/or synaptogyrin do play a role in regulating synaptic vesicle regeneration via a clathrin-­‐independent mechanism, it appears as though clathrin-­‐
mediated endocytosis may be partially compensating for the loss of those proteins. 46 Under conditions of increased neuronal activity (e.g., in dark-­‐adapted retinal rod photoreceptors), the clathrin-­‐mediated endocytosis pathway would be insufficient, thereby increasing the phenotypic severity. However, as previously discussed, the existence of such a clathrin-­‐independent endocytic mechanism is debated, and clathrin-­‐
mediated endocytosis appears to account for the vast majority of synaptic vesicle endocytosis at most synapses (Granseth et al., 2006; Heerssen et al., 2008). Synaptogyrin: a schizophrenia susceptibility gene? Schizophrenia is a complex mental disorder with a wide array of manifestations including hallucinations, delusions, and decreased emotional expression (Andreasen, 1995). While the exact cause of schizophrenia is unknown, there is significant evidence suggesting that there is a strong genetic component to the disorder with multiple loci contributing to disease susceptibility (Sullivan et al., 2003; Gejman et al., 2010). Several lines of evidence have implicated synaptogyrin as a candidate schizophrenia susceptibility gene. The chromosomal locus encoding human synaptogyrin 1 (22q11-­‐
13) has been linked with schizophrenia in a variety of genetic studies (Coon et al., 1994; Pulver et al., 1994; Gill et al., 1996; Verma et al., 2005; Bassett and Chow, 2008). Microarray analysis also identified synaptogyrin as one of the downregulated genes in a subset of schizophrenic patients whose prefrontal cortices were examined postmortem (Mirnics et al., 2000). Furthermore, several families in India and China with a history of schizophrenia and/or bipolar disorder have been found to have mutations within the synaptogyrin 1 locus (Verma et al., 2004; Cheng and Chen, 2007), and an association study in Italy identified several novel polymorphisms within the synaptogyrin locus, including a mutation that eliminates a potential serine phosphorylation site 47 (Iatropoulos et al., 2009). While definitive evidence linking synaptogyrin mutations to schizophrenia predisposition remains to be determined, synaptogyrin remains an intriguing candidate due to the changes in synaptic plasticity seen in knockout mice (Janz et al., 1999). The Drosophila larval NMJ as a model of synaptic development and function The neuromuscular junction (NMJ) of the Drosophila third instar larva has long been used as a system for studying synaptic transmission due to its easy accessibility, precise innervation patterns, and relative simplicity (Keshishian et al., 1996; Hoang and Chiba, 2001). Moreover, most of the key genes involved in synapse development, synaptic vesicle exo-­‐endocytosis, and intracellular trafficking are conserved in the Drosophila genome (Littleton, 2000; Lloyd et al., 2000). This system also provides access to a variety of genetic tools, most notably the GAL4/UAS system, which allows for spatially-­‐ and temporally-­‐targeted transgene expression (Brand and Perrimon, 1993; McGuire et al., 2004). Furthermore, the larval NMJ is similar to vertebrate central synapses in that it is glutamatergic, has graded responses to stimulation, and displays several forms of synaptic plasticity, including facilitation and post-­‐tetanic potentiation (Jan and Jan, 1976; Zhong and Wu, 1991; Broadie et al., 1997). These factors, combined with a short life cycle and easy maintenance, make Drosophila extremely amenable to study the genetic components of neurotransmission. 48 References Abmayr SM, Zhuang S, Geisbrecht ER (2008) Myoblast fusion in Drosophila. Methods Mol Biol 475:75-­‐97. Abraham C, Bai L, Leube RE (2011) Synaptogyrin-­‐dependent modulation of synaptic neurotransmission in Caenorhabditis elegans. Neuroscience 190:75-­‐88. Abraham C, Hutter H, Palfreyman MT, Spatkowski G, Weimer RM, Windoffer R, Jorgensen EM, Leube RE (2006) Synaptic tetraspan vesicle membrane proteins are conserved but not needed for synaptogenesis and neuronal function in Caenorhabditis elegans. Proc Natl Acad Sci U S A 103:8227-­‐8232. Ahle S, Ungewickell E (1986) Purification and properties of a new clathrin assembly protein. EMBO J 5:3143-­‐3149. Akbergenova Y, Bykhovskaia M (2009) Enhancement of the endosomal endocytic pathway increases quantal size. Mol Cell Neurosci 40:199-­‐206. Alder J, Xie ZP, Valtorta F, Greengard P, Poo M (1992a) Antibodies to synaptophysin interfere with transmitter secretion at neuromuscular synapses. Neuron 9:759-­‐
768. Alder J, Lu B, Valtorta F, Greengard P, Poo MM (1992b) Calcium-­‐dependent transmitter secretion reconstituted in Xenopus oocytes: requirement for synaptophysin. Science 257:657-­‐661. Alder J, Kanki H, Valtorta F, Greengard P, Poo MM (1995) Overexpression of synaptophysin enhances neurotransmitter secretion at Xenopus neuromuscular synapses. J Neurosci 15:511-­‐519. Ales E, Tabares L, Poyato JM, Valero V, Lindau M, Alvarez de Toledo G (1999) High calcium concentrations shift the mode of exocytosis to the kiss-­‐and-­‐run mechanism. Nat Cell Biol 1:40-­‐44. Andreasen NC (1995) Symptoms, signs, and diagnosis of schizophrenia. Lancet 346:477-­‐481. Arac D, Chen X, Khant HA, Ubach J, Ludtke SJ, Kikkawa M, Johnson AE, Chiu W, Sudhof TC, Rizo J (2006) Close membrane-­‐membrane proximity induced by Ca(2+)-­‐
dependent multivalent binding of synaptotagmin-­‐1 to phospholipids. Nat Struct Mol Biol 13:209-­‐217. Arthur CP, Stowell MH (2007) Structure of synaptophysin: a hexameric MARVEL-­‐
domain channel protein. Structure 15:707-­‐714. Barnekow A, Jahn R, Schartl M (1990) Synaptophysin: a substrate for the protein tyrosine kinase pp60c-­‐src in intact synaptic vesicles. Oncogene 5:1019-­‐1024. Bashkirov PV, Akimov SA, Evseev AI, Schmid SL, Zimmerberg J, Frolov VA (2008) GTPase cycle of dynamin is coupled to membrane squeeze and release, leading to spontaneous fission. Cell 135:1276-­‐1286. Bassett AS, Chow EW (2008) Schizophrenia and 22q11.2 deletion syndrome. Curr Psychiatry Rep 10:148-­‐157. Baumert M, Takei K, Hartinger J, Burger PM, Fischer von Mollard G, Maycox PR, De Camilli P, Jahn R (1990) P29: a novel tyrosine-­‐phosphorylated membrane protein present in small clear vesicles of neurons and endocrine cells. J Cell Biol 110:1285-­‐1294. 49 Belfort GM, Kandror KV (2003) Cellugyrin and synaptogyrin facilitate targeting of synaptophysin to a ubiquitous synaptic vesicle-­‐sized compartment in PC12 cells. J Biol Chem 278:47971-­‐47978. Belfort GM, Bakirtzi K, Kandror KV (2005) Cellugyrin induces biogenesis of synaptic-­‐
like microvesicles in PC12 cells. J Biol Chem 280:7262-­‐7272. Belizaire R, Komanduri C, Wooten K, Chen M, Thaller C, Janz R (2004) Characterization of synaptogyrin 3 as a new synaptic vesicle protein. J Comp Neurol 470:266-­‐281. Bennett MK, Calakos N, Scheller RH (1992) Syntaxin: a synaptic protein implicated in docking of synaptic vesicles at presynaptic active zones. Science 257:255-­‐259. Blondeau F, Ritter B, Allaire PD, Wasiak S, Girard M, Hussain NK, Angers A, Legendre-­‐
Guillemin V, Roy L, Boismenu D, Kearney RE, Bell AW, Bergeron JJ, McPherson PS (2004) Tandem MS analysis of brain clathrin-­‐coated vesicles reveals their critical involvement in synaptic vesicle recycling. Proc Natl Acad Sci U S A 101:3833-­‐3838. Bosse F, Hasse B, Pippirs U, Greiner-­‐Petter R, Muller HW (2003) Proteolipid plasmolipin: localization in polarized cells, regulated expression and lipid raft association in CNS and PNS myelin. J Neurochem 86:508-­‐518. Bowen ME, Weninger K, Brunger AT, Chu S (2004) Single molecule observation of liposome-­‐bilayer fusion thermally induced by soluble N-­‐ethyl maleimide sensitive-­‐factor attachment protein receptors (SNAREs). Biophys J 87:3569-­‐
3584. Bracher A, Kadlec J, Betz H, Weissenhorn W (2002) X-­‐ray structure of a neuronal complexin-­‐SNARE complex from squid. J Biol Chem 277:26517-­‐26523. Brand AH, Perrimon N (1993) Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118:401-­‐415. Brand SH, Laurie SM, Mixon MB, Castle JD (1991) Secretory carrier membrane proteins 31-­‐35 define a common protein composition among secretory carrier membranes. J Biol Chem 266:18949-­‐18957. Breckenridge LJ, Almers W (1987) Currents through the fusion pore that forms during exocytosis of a secretory vesicle. Nature 328:814-­‐817. Breckenridge WC, Morgan IG, Zanetta JP, Vincendon G (1973) Adult rat brain synaptic vesicles. II. Lipid composition. Biochim Biophys Acta 320:681-­‐686. Broadie K, Rushton E, Skoulakis EM, Davis RL (1997) Leonardo, a Drosophila 14-­‐3-­‐3 protein involved in learning, regulates presynaptic function. Neuron 19:391-­‐402. Brock LG, Coombs JS, Eccles JC (1952) The recording of potentials from motoneurones with an intracellular electrode. J Physiol 117:431-­‐460. Brodin L, Low P, Shupliakov O (2000) Sequential steps in clathrin-­‐mediated synaptic vesicle endocytosis. Curr Opin Neurobiol 10:312-­‐320. Buckley KM, Floor E, Kelly RB (1987) Cloning and sequence analysis of cDNA encoding p38, a major synaptic vesicle protein. J Cell Biol 105:2447-­‐2456. Burre J, Beckhaus T, Schagger H, Corvey C, Hofmann S, Karas M, Zimmermann H, Volknandt W (2006) Analysis of the synaptic vesicle proteome using three gel-­‐
based protein separation techniques. Proteomics 6:6250-­‐6262. Calakos N, Scheller RH (1994) Vesicle-­‐associated membrane protein and synaptophysin are associated on the synaptic vesicle. J Biol Chem 269:24534-­‐24537. 50 Cameron RS, Cameron PL, Castle JD (1986) A common spectrum of polypeptides occurs in secretion granule membranes of different exocrine glands. J Cell Biol 103:1299-­‐1313. Carrion-­‐Vazquez M, Fernandez AM, Chowen J, Nieto-­‐Sampedro M (1998) Brain Ac39/physophilin: cloning, coexpression and colocalization with synaptophysin. Eur J Neurosci 10:1153-­‐1166. Ceccarelli B, Hurlbut WP, Mauro A (1973) Turnover of transmitter and synaptic vesicles at the frog neuromuscular junction. J Cell Biol 57:499-­‐524. Chapman ER, Hanson PI, An S, Jahn R (1995) Ca2+ regulates the interaction between synaptotagmin and syntaxin 1. J Biol Chem 270:23667-­‐23671. Chen X, Arac D, Wang TM, Gilpin CJ, Zimmerberg J, Rizo J (2006) SNARE-­‐mediated lipid mixing depends on the physical state of the vesicles. Biophys J 90:2062-­‐2074. Chen X, Tomchick DR, Kovrigin E, Arac D, Machius M, Sudhof TC, Rizo J (2002) Three-­‐
dimensional structure of the complexin/SNARE complex. Neuron 33:397-­‐409. Cheng MC, Chen CH (2007) Identification of rare mutations of synaptogyrin 1 gene in patients with schizophrenia. J Psychiatr Res 41:1027-­‐1031. Cheong KH, Zacchetti D, Schneeberger EE, Simons K (1999) VIP17/MAL, a lipid raft-­‐
associated protein, is involved in apical transport in MDCK cells. Proc Natl Acad Sci U S A 96:6241-­‐6248. Cheung G, Jupp OJ, Cousin MA (2010) Activity-­‐dependent bulk endocytosis and clathrin-­‐
dependent endocytosis replenish specific synaptic vesicle pools in central nerve terminals. J Neurosci 30:8151-­‐8161. Chi S, Cao H, Chen J, McNiven MA (2008) Eps15 mediates vesicle trafficking from the trans-­‐Golgi network via an interaction with the clathrin adaptor AP-­‐1. Mol Biol Cell 19:3564-­‐3575. Cho RW, Song Y, Littleton JT (2010) Comparative analysis of Drosophila and mammalian complexins as fusion clamps and facilitators of neurotransmitter release. Mol Cell Neurosci 45:389-­‐397. Clayton EL, Cousin MA (2008) Differential labelling of bulk endocytosis in nerve terminals by FM dyes. Neurochem Int 53:51-­‐55. Clayton EL, Evans GJ, Cousin MA (2008) Bulk synaptic vesicle endocytosis is rapidly triggered during strong stimulation. J Neurosci 28:6627-­‐6632. Clayton EL, Anggono V, Smillie KJ, Chau N, Robinson PJ, Cousin MA (2009) The phospho-­‐
dependent dynamin-­‐syndapin interaction triggers activity-­‐dependent bulk endocytosis of synaptic vesicles. J Neurosci 29:7706-­‐7717. Cochilla AJ, Angleson JK, Betz WJ (1999) Monitoring secretory membrane with FM1-­‐43 fluorescence. Annu Rev Neurosci 22:1-­‐10. Coon H, Jensen S, Holik J, Hoff M, Myles-­‐Worsley M, Reimherr F, Wender P, Waldo M, Freedman R, Leppert M, et al. (1994) Genomic scan for genes predisposing to schizophrenia. Am J Med Genet 54:59-­‐71. Cousin MA (2009) Activity-­‐dependent bulk synaptic vesicle endocytosis-­‐-­‐a fast, high capacity membrane retrieval mechanism. Mol Neurobiol 39:185-­‐189. Cremona O, Di Paolo G, Wenk MR, Luthi A, Kim WT, Takei K, Daniell L, Nemoto Y, Shears SB, Flavell RA, McCormick DA, De Camilli P (1999) Essential role of phosphoinositide metabolism in synaptic vesicle recycling. Cell 99:179-­‐188. 51 Daly C, Ziff EB (2002) Ca2+-­‐dependent formation of a dynamin-­‐synaptophysin complex: potential role in synaptic vesicle endocytosis. J Biol Chem 277:9010-­‐9015. Daly C, Sugimori M, Moreira JE, Ziff EB, Llinas R (2000) Synaptophysin regulates clathrin-­‐independent endocytosis of synaptic vesicles. Proc Natl Acad Sci U S A 97:6120-­‐6125. de Beer T, Carter RE, Lobel-­‐Rice KE, Sorkin A, Overduin M (1998) Structure and Asn-­‐
Pro-­‐Phe binding pocket of the Eps15 homology domain. Science 281:1357-­‐1360. de Lange RP, de Roos AD, Borst JG (2003) Two modes of vesicle recycling in the rat calyx of Held. J Neurosci 23:10164-­‐10173. de Marco MC, Kremer L, Albar JP, Martinez-­‐Menarguez JA, Ballesta J, Garcia-­‐Lopez MA, Marazuela M, Puertollano R, Alonso MA (2001) BENE, a novel raft-­‐associated protein of the MAL proteolipid family, interacts with caveolin-­‐1 in human endothelial-­‐like ECV304 cells. J Biol Chem 276:23009-­‐23017. de Marco MC, Martin-­‐Belmonte F, Kremer L, Albar JP, Correas I, Vaerman JP, Marazuela M, Byrne JA, Alonso MA (2002) MAL2, a novel raft protein of the MAL family, is an essential component of the machinery for transcytosis in hepatoma HepG2 cells. J Cell Biol 159:37-­‐44. De Robertis E, Rodriguez De Lores Arnaiz G, Pellegrino De Iraldi A (1962) Isolation of synaptic vesicles from nerve endings of the rat brain. Nature 194:794-­‐795. De Robertis ED, Bennett HS (1955) Some features of the submicroscopic morphology of synapses in frog and earthworm. J Biophys Biochem Cytol 1:47-­‐58. Del Castillo J, Katz B (1954a) Quantal components of the end-­‐plate potential. J Physiol 124:560-­‐573. Del Castillo J, Katz B (1954b) Statistical factors involved in neuromuscular facilitation and depression. J Physiol 124:574-­‐585. Delgado R, Maureira C, Oliva C, Kidokoro Y, Labarca P (2000) Size of vesicle pools, rates of mobilization, and recycling at neuromuscular synapses of a Drosophila mutant, shibire. Neuron 28:941-­‐953. Dickman DK, Horne JA, Meinertzhagen IA, Schwarz TL (2005) A slowed classical pathway rather than kiss-­‐and-­‐run mediates endocytosis at synapses lacking synaptojanin and endophilin. Cell 123:521-­‐533. Dulubova I, Sugita S, Hill S, Hosaka M, Fernandez I, Sudhof TC, Rizo J (1999) A conformational switch in syntaxin during exocytosis: role of munc18. EMBO J 18:4372-­‐4382. Dunaevsky A, Connor EA (2000) F-­‐actin is concentrated in nonrelease domains at frog neuromuscular junctions. J Neurosci 20:6007-­‐6012. Eccles JC (1982) The synapse: from electrical to chemical transmission. Annu Rev Neurosci 5:325-­‐339. Edelmann L, Hanson PI, Chapman ER, Jahn R (1995) Synaptobrevin binding to synaptophysin: a potential mechanism for controlling the exocytotic fusion machine. EMBO J 14:224-­‐231. Elmqvist D, Quastel DM (1965) A quantitative study of end-­‐plate potentials in isolated human muscle. J Physiol 178:505-­‐529. Eshkind LG, Leube RE (1995) Mice lacking synaptophysin reproduce and form typical synaptic vesicles. Cell Tissue Res 282:423-­‐433. 52 Estrada B, Maeland AD, Gisselbrecht SS, Bloor JW, Brown NH, Michelson AM (2007) The MARVEL domain protein, Singles Bar, is required for progression past the pre-­‐
fusion complex stage of myoblast fusion. Dev Biol 307:328-­‐339. Evans GJ, Cousin MA (2005) Tyrosine phosphorylation of synaptophysin in synaptic vesicle recycling. Biochem Soc Trans 33:1350-­‐1353. Evans GJ, Cousin MA (2007) Activity-­‐dependent control of slow synaptic vesicle endocytosis by cyclin-­‐dependent kinase 5. J Neurosci 27:401-­‐411. Farsad K, Ringstad N, Takei K, Floyd SR, Rose K, De Camilli P (2001) Generation of high curvature membranes mediated by direct endophilin bilayer interactions. J Cell Biol 155:193-­‐200. Fatt P, Katz B (1951) An analysis of the end-­‐plate potential recorded with an intracellular electrode. J Physiol 115:320-­‐370. Fatt P, Katz B (1952) Spontaneous subthreshold activity at motor nerve endings. J Physiol 117:109-­‐128. Fernandez I, Ubach J, Dulubova I, Zhang X, Sudhof TC, Rizo J (1998) Three-­‐dimensional structure of an evolutionarily conserved N-­‐terminal domain of syntaxin 1A. Cell 94:841-­‐849. Fernandez I, Arac D, Ubach J, Gerber SH, Shin O, Gao Y, Anderson RG, Sudhof TC, Rizo J (2001) Three-­‐dimensional structure of the synaptotagmin 1 C2B-­‐domain: synaptotagmin 1 as a phospholipid binding machine. Neuron 32:1057-­‐1069. Fernandez JM, Neher E, Gomperts BD (1984) Capacitance measurements reveal stepwise fusion events in degranulating mast cells. Nature 312:453-­‐455. Fernandez-­‐Chacon R, Sudhof TC (1999) Genetics of synaptic vesicle function: toward the complete functional anatomy of an organelle. Annu Rev Physiol 61:753-­‐776. Fernandez-­‐Chacon R, Sudhof TC (2000) Novel SCAMPs lacking NPF repeats: ubiquitous and synaptic vesicle-­‐specific forms implicate SCAMPs in multiple membrane-­‐
trafficking functions. J Neurosci 20:7941-­‐7950. Fernandez-­‐Chacon R, Alvarez de Toledo G, Hammer RE, Sudhof TC (1999) Analysis of SCAMP1 function in secretory vesicle exocytosis by means of gene targeting in mice. J Biol Chem 274:32551-­‐32554. Fernandez-­‐Chacon R, Achiriloaie M, Janz R, Albanesi JP, Sudhof TC (2000) SCAMP1 function in endocytosis. J Biol Chem 275:12752-­‐12756. Fernandez-­‐Chacon R, Konigstorfer A, Gerber SH, Garcia J, Matos MF, Stevens CF, Brose N, Rizo J, Rosenmund C, Sudhof TC (2001) Synaptotagmin I functions as a calcium regulator of release probability. Nature 410:41-­‐49. Fesce R, Grohovaz F, Valtorta F, Meldolesi J (1994) Neurotransmitter release: fusion or 'kiss-­‐and-­‐run'? Trends Cell Biol 4:1-­‐4. Fleishman SJ, Unger VM, Yeager M, Ben-­‐Tal N (2004) A Calpha model for the transmembrane alpha helices of gap junction intercellular channels. Mol Cell 15:879-­‐888. Franzini-­‐Armstrong C, Jorgensen AO (1994) Structure and development of E-­‐C coupling units in skeletal muscle. Annu Rev Physiol 56:509-­‐534. Furuse M, Tsukita S (2006) Claudins in occluding junctions of humans and flies. Trends Cell Biol 16:181-­‐188. Furuse M, Fujimoto K, Sato N, Hirase T, Tsukita S (1996) Overexpression of occludin, a tight junction-­‐associated integral membrane protein, induces the formation of 53 intracellular multilamellar bodies bearing tight junction-­‐like structures. J Cell Sci 109 ( Pt 2):429-­‐435. Furuse M, Hirase T, Itoh M, Nagafuchi A, Yonemura S, Tsukita S (1993) Occludin: a novel integral membrane protein localizing at tight junctions. J Cell Biol 123:1777-­‐
1788. Fykse EM, Takei K, Walch-­‐Solimena C, Geppert M, Jahn R, De Camilli P, Sudhof TC (1993) Relative properties and localizations of synaptic vesicle protein isoforms: the case of the synaptophysins. J Neurosci 13:4997-­‐5007. Gad H, Low P, Zotova E, Brodin L, Shupliakov O (1998) Dissociation between Ca2+-­‐
triggered synaptic vesicle exocytosis and clathrin-­‐mediated endocytosis at a central synapse. Neuron 21:607-­‐616. Gaidarov I, Keen JH (1999) Phosphoinositide-­‐AP-­‐2 interactions required for targeting to plasma membrane clathrin-­‐coated pits. J Cell Biol 146:755-­‐764. Galbiati F, Volonte D, Engelman JA, Scherer PE, Lisanti MP (1999) Targeted down-­‐
regulation of caveolin-­‐3 is sufficient to inhibit myotube formation in differentiating C2C12 myoblasts. Transient activation of p38 mitogen-­‐activated protein kinase is required for induction of caveolin-­‐3 expression and subsequent myotube formation. J Biol Chem 274:30315-­‐30321. Galli T, McPherson PS, De Camilli P (1996) The V0 sector of the V-­‐ATPase, synaptobrevin, and synaptophysin are associated on synaptic vesicles in a Triton X-­‐100-­‐resistant, freeze-­‐thawing sensitive, complex. J Biol Chem 271:2193-­‐2198. Gallop JL, Butler PJ, McMahon HT (2005) Endophilin and CtBP/BARS are not acyl transferases in endocytosis or Golgi fission. Nature 438:675-­‐678. Gallwitz D, Jahn R (2003) The riddle of the Sec1/Munc-­‐18 proteins -­‐ new twists added to their interactions with SNAREs. Trends Biochem Sci 28:113-­‐116. Gejman PV, Sanders AR, Duan J (2010) The role of genetics in the etiology of schizophrenia. Psychiatr Clin North Am 33:35-­‐66. Geppert M, Goda Y, Hammer RE, Li C, Rosahl TW, Stevens CF, Sudhof TC (1994) Synaptotagmin I: a major Ca2+ sensor for transmitter release at a central synapse. Cell 79:717-­‐727. Gill M et al. (1996) A combined analysis of D22S278 marker alleles in affected sib-­‐pairs: support for a susceptibility locus for schizophrenia at chromosome 22q12. Schizophrenia Collaborative Linkage Group (Chromosome 22). Am J Med Genet 67:40-­‐45. Gincel D, Shoshan-­‐Barmatz V (2002) The synaptic vesicle protein synaptophysin: purification and characterization of its channel activity. Biophys J 83:3223-­‐3229. Giraudo CG, Eng WS, Melia TJ, Rothman JE (2006) A clamping mechanism involved in SNARE-­‐dependent exocytosis. Science 313:676-­‐680. Gordon SL, Leube RE, Cousin MA (2011) Synaptophysin Is Required for Synaptobrevin Retrieval during Synaptic Vesicle Endocytosis. J Neurosci 31:14032-­‐14036. Granseth B, Odermatt B, Royle SJ, Lagnado L (2006) Clathrin-­‐mediated endocytosis is the dominant mechanism of vesicle retrieval at hippocampal synapses. Neuron 51:773-­‐786. Gray EG, Whittaker VP (1962) The isolation of nerve endings from brain: an electron-­‐
microscopic study of cell fragments derived by homogenization and centrifugation. J Anat 96:79-­‐88. 54 Grigliatti TA, Hall L, Rosenbluth R, Suzuki DT (1973) Temperature-­‐sensitive mutations in Drosophila melanogaster. XIV. A selection of immobile adults. Mol Gen Genet 120:107-­‐114. Guo Z, Liu L, Cafiso D, Castle D (2002) Perturbation of a very late step of regulated exocytosis by a secretory carrier membrane protein (SCAMP2)-­‐derived peptide. J Biol Chem 277:35357-­‐35363. Haass NK, Kartenbeck MA, Leube RE (1996) Pantophysin is a ubiquitously expressed synaptophysin homologue and defines constitutive transport vesicles. J Cell Biol 134:731-­‐746. Haffner C, Takei K, Chen H, Ringstad N, Hudson A, Butler MH, Salcini AE, Di Fiore PP, De Camilli P (1997) Synaptojanin 1: localization on coated endocytic intermediates in nerve terminals and interaction of its 170 kDa isoform with Eps15. FEBS Lett 419:175-­‐180. Hamacher M, Pippirs U, Kohler A, Muller HW, Bosse F (2001) Plasmolipin: genomic structure, chromosomal localization, protein expression pattern, and putative association with Bardet-­‐Biedl syndrome. Mamm Genome 12:933-­‐937. Hanson PI, Roth R, Morisaki H, Jahn R, Heuser JE (1997) Structure and conformational changes in NSF and its membrane receptor complexes visualized by quick-­‐
freeze/deep-­‐etch electron microscopy. Cell 90:523-­‐535. Hao W, Tan Z, Prasad K, Reddy KK, Chen J, Prestwich GD, Falck JR, Shears SB, Lafer EM (1997) Regulation of AP-­‐3 function by inositides. Identification of phosphatidylinositol 3,4,5-­‐trisphosphate as a potent ligand. J Biol Chem 272:6393-­‐6398. Harata N, Ryan TA, Smith SJ, Buchanan J, Tsien RW (2001) Visualizing recycling synaptic vesicles in hippocampal neurons by FM 1-­‐43 photoconversion. Proc Natl Acad Sci U S A 98:12748-­‐12753. Harris TW, Hartwieg E, Horvitz HR, Jorgensen EM (2000) Mutations in synaptojanin disrupt synaptic vesicle recycling. J Cell Biol 150:589-­‐600. Hata Y, Slaughter CA, Sudhof TC (1993) Synaptic vesicle fusion complex contains unc-­‐18 homologue bound to syntaxin. Nature 366:347-­‐351. Haucke V, De Camilli P (1999) AP-­‐2 recruitment to synaptotagmin stimulated by tyrosine-­‐based endocytic motifs. Science 285:1268-­‐1271. He L, Wu LG (2007) The debate on the kiss-­‐and-­‐run fusion at synapses. Trends Neurosci 30:447-­‐455. Heerssen H, Fetter RD, Davis GW (2008) Clathrin dependence of synaptic-­‐vesicle formation at the Drosophila neuromuscular junction. Curr Biol 18:401-­‐409. Heidelberger R, Heinemann C, Neher E, Matthews G (1994) Calcium dependence of the rate of exocytosis in a synaptic terminal. Nature 371:513-­‐515. Heuser JE, Reese TS (1973) Evidence for recycling of synaptic vesicle membrane during transmitter release at the frog neuromuscular junction. J Cell Biol 57:315-­‐344. Heuser JE, Reese TS, Landis DMD (1974) Functional changes in frog neuromuscular junctions studied with freeze-­‐fracture. Journal of Neurocytology 3:109-­‐131. Heuser JE, Reese TS, Dennis MJ, Jan Y, Jan L, Evans L (1979) Synaptic vesicle exocytosis captured by quick freezing and correlated with quantal transmitter release. J Cell Biol 81:275-­‐300. 55 Hinshaw JE (2000) Dynamin and its role in membrane fission. Annu Rev Cell Dev Biol 16:483-­‐519. Hirst J, Robinson MS (1998) Clathrin and adaptors. Biochim Biophys Acta 1404:173-­‐
193. Hoang B, Chiba A (2001) Single-­‐cell analysis of Drosophila larval neuromuscular synapses. Dev Biol 229:55-­‐70. Hubbard C, Singleton D, Rauch M, Jayasinghe S, Cafiso D, Castle D (2000) The secretory carrier membrane protein family: structure and membrane topology. Mol Biol Cell 11:2933-­‐2947. Hubner K, Windoffer R, Hutter H, Leube RE (2002) Tetraspan vesicle membrane proteins: synthesis, subcellular localization, and functional properties. Int Rev Cytol 214:103-­‐159. Huntwork S, Littleton JT (2007) A complexin fusion clamp regulates spontaneous neurotransmitter release and synaptic growth. Nat Neurosci 10:1235-­‐1237. Hussain NK, Yamabhai M, Ramjaun AR, Guy AM, Baranes D, O'Bryan JP, Der CJ, Kay BK, McPherson PS (1999) Splice variants of intersectin are components of the endocytic machinery in neurons and nonneuronal cells. J Biol Chem 274:15671-­‐
15677. Iatropoulos P, Gardella R, Valsecchi P, Magri C, Ratti C, Podavini D, Rossi G, Gennarelli M, Sacchetti E, Barlati S (2009) Association study and mutational screening of SYNGR1 as a candidate susceptibility gene for schizophrenia. Psychiatr Genet 19:237-­‐243. Jahn R, Sudhof TC (1994) Synaptic vesicles and exocytosis. Annu Rev Neurosci 17:219-­‐
246. Jahn R, Schiebler W, Ouimet C, Greengard P (1985) A 38,000-­‐dalton membrane protein (p38) present in synaptic vesicles. Proc Natl Acad Sci U S A 82:4137-­‐4141. Jan LY, Jan YN (1976) L-­‐glutamate as an excitatory transmitter at the Drosophila larval neuromuscular junction. J Physiol 262:215-­‐236. Janz R, Sudhof TC (1998) Cellugyrin, a novel ubiquitous form of synaptogyrin that is phosphorylated by pp60c-­‐src. J Biol Chem 273:2851-­‐2857. Janz R, Sudhof TC, Hammer RE, Unni V, Siegelbaum SA, Bolshakov VY (1999) Essential roles in synaptic plasticity for synaptogyrin I and synaptophysin I. Neuron 24:687-­‐700. Johnston PA, Sudhof TC (1990) The multisubunit structure of synaptophysin. Relationship between disulfide bonding and homo-­‐oligomerization. J Biol Chem 265:8869-­‐8873. Kedra D, Pan HQ, Seroussi E, Fransson I, Guilbaud C, Collins JE, Dunham I, Blennow E, Roe BA, Piehl F, Dumanski JP (1998) Characterization of the human synaptogyrin gene family. Hum Genet 103:131-­‐141. Keshishian H, Broadie K, Chiba A, Bate M (1996) The drosophila neuromuscular junction: a model system for studying synaptic development and function. Annu Rev Neurosci 19:545-­‐575. Knaus P, Marqueze-­‐Pouey B, Scherer H, Betz H (1990) Synaptoporin, a novel putative channel protein of synaptic vesicles. Neuron 5:453-­‐462. Koenig JH, Ikeda K (1983) Evidence for a presynaptic blockage of transmission in a temperature-­‐sensitive mutant of Drosophila. J Neurobiol 14:411-­‐419. 56 Koenig JH, Ikeda K (1989) Disappearance and reformation of synaptic vesicle membrane upon transmitter release observed under reversible blockage of membrane retrieval. J Neurosci 9:3844-­‐3860. Komazaki S, Nishi M, Kangawa K, Takeshima H (1999) Immunolocalization of mitsugumin29 in developing skeletal muscle and effects of the protein expressed in amphibian embryonic cells. Dev Dyn 215:87-­‐95. Komazaki S, Nishi M, Takeshima H, Nakamura H (2001) Abnormal formation of sarcoplasmic reticulum networks and triads during early development of skeletal muscle cells in mitsugumin29-­‐deficient mice. Dev Growth Differ 43:717-­‐
723. Kronengold J, Trexler EB, Bukauskas FF, Bargiello TA, Verselis VK (2003) Single-­‐
channel SCAM identifies pore-­‐lining residues in the first extracellular loop and first transmembrane domains of Cx46 hemichannels. J Gen Physiol 122:389-­‐405. Kuffler SW (1942a) Further study on transmission in an isolated nerve-­‐muscle fibre preparation. J Neurophysiol 5:309-­‐322. Kuffler SW (1942b) Electric potential changes at an isolated nerve-­‐muscle junction. J Neurophysiol 5:18-­‐26. Kumashiro S, Lu YF, Tomizawa K, Matsushita M, Wei FY, Matsui H (2005) Regulation of synaptic vesicle recycling by calcineurin in different vesicle pools. Neurosci Res 51:435-­‐443. Kuromi H, Kidokoro Y (1998) Two distinct pools of synaptic vesicles in single presynaptic boutons in a temperature-­‐sensitive Drosophila mutant, shibire. Neuron 20:917-­‐925. Kuromi H, Kidokoro Y (2000) Tetanic stimulation recruits vesicles from reserve pool via a cAMP-­‐mediated process in Drosophila synapses. Neuron 27:133-­‐143. Kuromi H, Kidokoro Y (2002) Selective replenishment of two vesicle pools depends on the source of Ca2+ at the Drosophila synapse. Neuron 35:333-­‐343. Kuromi H, Kidokoro Y (2003) Two synaptic vesicle pools, vesicle recruitment and replenishment of pools at the Drosophila neuromuscular junction. J Neurocytol 32:551-­‐565. Kwon SE, Chapman ER (2011) Synaptophysin regulates the kinetics of synaptic vesicle endocytosis in central neurons. Neuron 70:847-­‐854. Laurie SM, Cain CC, Lienhard GE, Castle JD (1993) The glucose transporter GluT4 and secretory carrier membrane proteins (SCAMPs) colocalize in rat adipocytes and partially segregate during insulin stimulation. J Biol Chem 268:19110-­‐19117. Leube RE (1994) Expression of the synaptophysin gene family is not restricted to neuronal and neuroendocrine differentiation in rat and human. Differentiation 56:163-­‐171. Leube RE (1995) The topogenic fate of the polytopic transmembrane proteins, synaptophysin and connexin, is determined by their membrane-­‐spanning domains. J Cell Sci 108 ( Pt 3):883-­‐894. Leube RE, Kaiser P, Seiter A, Zimbelmann R, Franke WW, Rehm H, Knaus P, Prior P, Betz H, Reinke H, et al. (1987) Synaptophysin: molecular organization and mRNA expression as determined from cloned cDNA. EMBO J 6:3261-­‐3268. 57 Li C, Ullrich B, Zhang JZ, Anderson RG, Brose N, Sudhof TC (1995) Ca(2+)-­‐dependent and -­‐independent activities of neural and non-­‐neural synaptotagmins. Nature 375:594-­‐599. Li F, Pincet F, Perez E, Eng WS, Melia TJ, Rothman JE, Tareste D (2007) Energetics and dynamics of SNAREpin folding across lipid bilayers. Nat Struct Mol Biol 14:890-­‐
896. Linstedt AD, Vetter ML, Bishop JM, Kelly RB (1992) Specific association of the proto-­‐
oncogene product pp60c-­‐src with an intracellular organelle, the PC12 synaptic vesicle. J Cell Biol 117:1077-­‐1084. Littleton JT (2000) A genomic analysis of membrane trafficking and neurotransmitter release in Drosophila. J Cell Biol 150:F77-­‐82. Littleton JT, Stern M, Schulze K, Perin M, Bellen HJ (1993) Mutational analysis of Drosophila synaptotagmin demonstrates its essential role in Ca(2+)-­‐activated neurotransmitter release. Cell 74:1125-­‐1134. Liu L, Guo Z, Tieu Q, Castle A, Castle D (2002) Role of secretory carrier membrane protein SCAMP2 in granule exocytosis. Mol Biol Cell 13:4266-­‐4278. Llinas R, Sugimori M, Silver RB (1992) Microdomains of high calcium concentration in a presynaptic terminal. Science 256:677-­‐679. Lloyd TE, Verstreken P, Ostrin EJ, Phillippi A, Lichtarge O, Bellen HJ (2000) A genome-­‐
wide search for synaptic vesicle cycle proteins in Drosophila. Neuron 26:45-­‐50. Lowe AW, Madeddu L, Kelly RB (1988) Endocrine secretory granules and neuronal synaptic vesicles have three integral membrane proteins in common. J Cell Biol 106:51-­‐59. Maeda S, Nakagawa S, Suga M, Yamashita E, Oshima A, Fujiyoshi Y, Tsukihara T (2009) Structure of the connexin 26 gap junction channel at 3.5 A resolution. Nature 458:597-­‐602. Magal LG, Yaffe Y, Shepshelovich J, Aranda JF, de Marco Mdel C, Gaus K, Alonso MA, Hirschberg K (2009) Clustering and lateral concentration of raft lipids by the MAL protein. Mol Biol Cell 20:3751-­‐3762. Masuda M, Takeda S, Sone M, Ohki T, Mori H, Kamioka Y, Mochizuki N (2006) Endophilin BAR domain drives membrane curvature by two newly identified structure-­‐based mechanisms. EMBO J 25:2889-­‐2897. Maximov A, Tang J, Yang X, Pang ZP, Sudhof TC (2009) Complexin controls the force transfer from SNARE complexes to membranes in fusion. Science 323:516-­‐521. Maycox PR, Deckwerth T, Hell JW, Jahn R (1988) Glutamate uptake by brain synaptic vesicles. Energy dependence of transport and functional reconstitution in proteoliposomes. J Biol Chem 263:15423-­‐15428. McGuire SE, Roman G, Davis RL (2004) Gene expression systems in Drosophila: a synthesis of time and space. Trends Genet 20:384-­‐391. McMahon HT, Missler M, Li C, Sudhof TC (1995) Complexins: cytosolic proteins that regulate SNAP receptor function. Cell 83:111-­‐119. McMahon HT, Bolshakov VY, Janz R, Hammer RE, Siegelbaum SA, Sudhof TC (1996) Synaptophysin, a major synaptic vesicle protein, is not essential for neurotransmitter release. Proc Natl Acad Sci U S A 93:4760-­‐4764. 58 McPherson PS, Garcia EP, Slepnev VI, David C, Zhang X, Grabs D, Sossin WS, Bauerfeind R, Nemoto Y, De Camilli P (1996) A presynaptic inositol-­‐5-­‐phosphatase. Nature 379:353-­‐357. Miller TM, Heuser JE (1984) Endocytosis of synaptic vesicle membrane at the frog neuromuscular junction. J Cell Biol 98:685-­‐698. Mirnics K, Middleton FA, Marquez A, Lewis DA, Levitt P (2000) Molecular characterization of schizophrenia viewed by microarray analysis of gene expression in prefrontal cortex. Neuron 28:53-­‐67. Montecucco C, Schiavo G, Pantano S (2005) SNARE complexes and neuroexocytosis: how many, how close? Trends Biochem Sci 30:367-­‐372. Mullany PM, Lynch MA (1998) Evidence for a role for synaptophysin in expression of long-­‐term potentiation in rat dentate gyrus. Neuroreport 9:2489-­‐2494. Murthy VN, De Camilli P (2003) Cell biology of the presynaptic terminal. Annu Rev Neurosci 26:701-­‐728. Mutch SA, Kensel-­‐Hammes P, Gadd JC, Fujimoto BS, Allen RW, Schiro PG, Lorenz RM, Kuyper CL, Kuo JS, Bajjalieh SM, Chiu DT (2011) Protein quantification at the single vesicle level reveals that a subset of synaptic vesicle proteins are trafficked with high precision. J Neurosci 31:1461-­‐1470. Nagaraj RY, Nosek CM, Brotto MA, Nishi M, Takeshima H, Nosek TM, Ma J (2000) Increased susceptibility to fatigue of slow-­‐ and fast-­‐twitch muscles from mice lacking the MG29 gene. Physiol Genomics 4:43-­‐49. Navone F, Jahn R, Di Gioia G, Stukenbrok H, Greengard P, De Camilli P (1986) Protein p38: an integral membrane protein specific for small vesicles of neurons and neuroendocrine cells. J Cell Biol 103:2511-­‐2527. Nishi M, Komazaki S, Kurebayashi N, Ogawa Y, Noda T, Iino M, Takeshima H (1999) Abnormal features in skeletal muscle from mice lacking mitsugumin29. J Cell Biol 147:1473-­‐1480. Nonet ML (1999) Visualization of synaptic specializations in live C. elegans with synaptic vesicle protein-­‐GFP fusions. J Neurosci Methods 89:33-­‐40. Nonet ML, Holgado AM, Brewer F, Serpe CJ, Norbeck BA, Holleran J, Wei L, Hartwieg E, Jorgensen EM, Alfonso A (1999) UNC-­‐11, a Caenorhabditis elegans AP180 homologue, regulates the size and protein composition of synaptic vesicles. Mol Biol Cell 10:2343-­‐2360. Nusrat A, Parkos CA, Verkade P, Foley CS, Liang TW, Innis-­‐Whitehouse W, Eastburn KK, Madara JL (2000) Tight junctions are membrane microdomains. J Cell Sci 113 ( Pt 10):1771-­‐1781. Oberhauser AF, Fernandez JM (1996) A fusion pore phenotype in mast cells of the ruby-­‐
eye mouse. Proc Natl Acad Sci U S A 93:14349-­‐14354. Paillart C, Li J, Matthews G, Sterling P (2003) Endocytosis and vesicle recycling at a ribbon synapse. J Neurosci 23:4092-­‐4099. Palay SL, Palade GE (1955) The fine structure of neurons. J Biophys Biochem Cytol 1:69-­‐
88. Pang DT, Wang JK, Valtorta F, Benfenati F, Greengard P (1988) Protein tyrosine phosphorylation in synaptic vesicles. Proc Natl Acad Sci U S A 85:762-­‐766. 59 Pang ZP, Sun J, Rizo J, Maximov A, Sudhof TC (2006) Genetic analysis of synaptotagmin 2 in spontaneous and Ca2+-­‐triggered neurotransmitter release. EMBO J 25:2039-­‐
2050. Paoluzi S, Castagnoli L, Lauro I, Salcini AE, Coda L, Fre S, Confalonieri S, Pelicci PG, Di Fiore PP, Cesareni G (1998) Recognition specificity of individual EH domains of mammals and yeast. EMBO J 17:6541-­‐6550. Parton RG, Simons K (2007) The multiple faces of caveolae. Nat Rev Mol Cell Biol 8:185-­‐
194. Pennuto M, Bonanomi D, Benfenati F, Valtorta F (2003) Synaptophysin I controls the targeting of VAMP2/synaptobrevin II to synaptic vesicles. Mol Biol Cell 14:4909-­‐
4919. Pennuto M, Dunlap D, Contestabile A, Benfenati F, Valtorta F (2002) Fluorescence resonance energy transfer detection of synaptophysin I and vesicle-­‐associated membrane protein 2 interactions during exocytosis from single live synapses. Mol Biol Cell 13:2706-­‐2717. Perin MS, Fried VA, Mignery GA, Jahn R, Sudhof TC (1990) Phospholipid binding by a synaptic vesicle protein homologous to the regulatory region of protein kinase C. Nature 345:260-­‐263. Puertollano R, Alonso MA (1999) MAL, an integral element of the apical sorting machinery, is an itinerant protein that cycles between the trans-­‐Golgi network and the plasma membrane. Mol Biol Cell 10:3435-­‐3447. Pulver AE, Karayiorgou M, Wolyniec PS, Lasseter VK, Kasch L, Nestadt G, Antonarakis S, Housman D, Kazazian HH, Meyers D, et al. (1994) Sequential strategy to identify a susceptibility gene for schizophrenia: report of potential linkage on chromosome 22q12-­‐q13.1: Part 1. Am J Med Genet 54:36-­‐43. Pumplin DW, Reese TS, Llinas R (1981) Are the presynaptic membrane particles the calcium channels? Proc Natl Acad Sci U S A 78:7210-­‐7213. Pyle JL, Kavalali ET, Piedras-­‐Renteria ES, Tsien RW (2000) Rapid reuse of readily releasable pool vesicles at hippocampal synapses. Neuron 28:221-­‐231. Raleigh DR, Marchiando AM, Zhang Y, Shen L, Sasaki H, Wang Y, Long M, Turner JR (2010) Tight junction-­‐associated MARVEL proteins marveld3, tricellulin, and occludin have distinct but overlapping functions. Mol Biol Cell 21:1200-­‐1213. Ramnarayanan SP, Tuma PL (2011) MAL, but not MAL2, expression promotes the formation of cholesterol-­‐dependent membrane domains that recruit apical proteins. Biochem J 439:497-­‐504. Rehm H, Wiedenmann B, Betz H (1986) Molecular characterization of synaptophysin, a major calcium-­‐binding protein of the synaptic vesicle membrane. EMBO J 5:535-­‐
541. Reim K, Mansour M, Varoqueaux F, McMahon HT, Sudhof TC, Brose N, Rosenmund C (2001) Complexins regulate a late step in Ca2+-­‐dependent neurotransmitter release. Cell 104:71-­‐81. Richards DA, Guatimosim C, Betz WJ (2000) Two endocytic recycling routes selectively fill two vesicle pools in frog motor nerve terminals. Neuron 27:551-­‐559. Richards DA, Rizzoli SO, Betz WJ (2004) Effects of wortmannin and latrunculin A on slow endocytosis at the frog neuromuscular junction. J Physiol 557:77-­‐91. 60 Richards DA, Guatimosim C, Rizzoli SO, Betz WJ (2003) Synaptic vesicle pools at the frog neuromuscular junction. Neuron 39:529-­‐541. Ringstad N, Nemoto Y, De Camilli P (1997) The SH3p4/Sh3p8/SH3p13 protein family: binding partners for synaptojanin and dynamin via a Grb2-­‐like Src homology 3 domain. Proc Natl Acad Sci U S A 94:8569-­‐8574. Rizo J, Chen X, Arac D (2006) Unraveling the mechanisms of synaptotagmin and SNARE function in neurotransmitter release. Trends Cell Biol 16:339-­‐350. Rizzoli SO, Betz WJ (2005) Synaptic vesicle pools. Nat Rev Neurosci 6:57-­‐69. Rizzoli SO, Jahn R (2007) Kiss-­‐and-­‐run, collapse and 'readily retrievable' vesicles. Traffic 8:1137-­‐1144. Roos J, Kelly RB (1999) The endocytic machinery in nerve terminals surrounds sites of exocytosis. Curr Biol 9:1411-­‐1414. Roux A, Uyhazi K, Frost A, De Camilli P (2006) GTP-­‐dependent twisting of dynamin implicates constriction and tension in membrane fission. Nature 441:528-­‐531. Rubenstein JL, Greengard P, Czernik AJ (1993) Calcium-­‐dependent serine phosphorylation of synaptophysin. Synapse 13:161-­‐172. Saitou M, Furuse M, Sasaki H, Schulzke JD, Fromm M, Takano H, Noda T, Tsukita S (2000) Complex phenotype of mice lacking occludin, a component of tight junction strands. Mol Biol Cell 11:4131-­‐4142. Sanchez-­‐Pulido L, Martin-­‐Belmonte F, Valencia A, Alonso MA (2002) MARVEL: a conserved domain involved in membrane apposition events. Trends Biochem Sci 27:599-­‐601. Sankaranarayanan S, Ryan TA (2000) Real-­‐time measurements of vesicle-­‐SNARE recycling in synapses of the central nervous system. Nat Cell Biol 2:197-­‐204. Schaub JR, Lu X, Doneske B, Shin YK, McNew JA (2006) Hemifusion arrest by complexin is relieved by Ca2+-­‐synaptotagmin I. Nat Struct Mol Biol 13:748-­‐750. Schikorski T, Stevens CF (1997) Quantitative ultrastructural analysis of hippocampal excitatory synapses. J Neurosci 17:5858-­‐5867. Schikorski T, Stevens CF (2001) Morphological correlates of functionally defined synaptic vesicle populations. Nat Neurosci 4:391-­‐395. Schmidt A, Wolde M, Thiele C, Fest W, Kratzin H, Podtelejnikov AV, Witke W, Huttner WB, Soling HD (1999) Endophilin I mediates synaptic vesicle formation by transfer of arachidonate to lysophosphatidic acid. Nature 401:133-­‐141. Schmitt U, Tanimoto N, Seeliger M, Schaeffel F, Leube RE (2009) Detection of behavioral alterations and learning deficits in mice lacking synaptophysin. Neuroscience 162:234-­‐243. Schuske KR, Richmond JE, Matthies DS, Davis WS, Runz S, Rube DA, van der Bliek AM, Jorgensen EM (2003) Endophilin is required for synaptic vesicle endocytosis by localizing synaptojanin. Neuron 40:749-­‐762. Shimuta M, Komazaki S, Nishi M, Iino M, Nakagawara K, Takeshima H (1998) Structure and expression of mitsugumin29 gene. FEBS Lett 431:263-­‐267. Shupliakov O (2009) The synaptic vesicle cluster: a source of endocytic proteins during neurotransmitter release. Neuroscience 158:204-­‐210. Siebert A, Lottspeich F, Nelson N, Betz H (1994) Purification of the synaptic vesicle-­‐
binding protein physophilin. Identification as 39-­‐kDa subunit of the vacuolar H(+)-­‐ATPase. J Biol Chem 269:28329-­‐28334. 61 Singleton DR, Wu TT, Castle JD (1997) Three mammalian SCAMPs (secretory carrier membrane proteins) are highly related products of distinct genes having similar subcellular distributions. J Cell Sci 110 ( Pt 17):2099-­‐2107. Skerrett IM, Aronowitz J, Shin JH, Cymes G, Kasperek E, Cao FL, Nicholson BJ (2002) Identification of amino acid residues lining the pore of a gap junction channel. J Cell Biol 159:349-­‐360. Slepnev VI, De Camilli P (2000) Accessory factors in clathrin-­‐dependent synaptic vesicle endocytosis. Nat Rev Neurosci 1:161-­‐172. Smith CJ, Pearse BM (1999) Clathrin: anatomy of a coat protein. Trends Cell Biol 9:335-­‐
338. Sollner T, Whiteheart SW, Brunner M, Erdjument-­‐Bromage H, Geromanos S, Tempst P, Rothman JE (1993) SNAP receptors implicated in vesicle targeting and fusion. Nature 362:318-­‐324. Song W, Zinsmaier KE (2003) Endophilin and synaptojanin hook up to promote synaptic vesicle endocytosis. Neuron 40:665-­‐667. Sorensen JB, Wiederhold K, Muller EM, Milosevic I, Nagy G, de Groot BL, Grubmuller H, Fasshauer D (2006) Sequential N-­‐ to C-­‐terminal SNARE complex assembly drives priming and fusion of secretory vesicles. EMBO J 25:955-­‐966. Spiwoks-­‐Becker I, Vollrath L, Seeliger MW, Jaissle G, Eshkind LG, Leube RE (2001) Synaptic vesicle alterations in rod photoreceptors of synaptophysin-­‐deficient mice. Neuroscience 107:127-­‐142. Stenius K, Janz R, Sudhof TC, Jahn R (1995) Structure of synaptogyrin (p29) defines novel synaptic vesicle protein. J Cell Biol 131:1801-­‐1809. Stevens CF, Williams JH (2000) "Kiss and run" exocytosis at hippocampal synapses. Proc Natl Acad Sci U S A 97:12828-­‐12833. Stowell MH, Marks B, Wigge P, McMahon HT (1999) Nucleotide-­‐dependent conformational changes in dynamin: evidence for a mechanochemical molecular spring. Nat Cell Biol 1:27-­‐32. Sudhof TC (2000) The synaptic vesicle cycle revisited. Neuron 28:317-­‐320. Sudhof TC (2004) The synaptic vesicle cycle. Annu Rev Neurosci 27:509-­‐547. Sudhof TC, Jahn R (1991) Proteins of synaptic vesicles involved in exocytosis and membrane recycling. Neuron 6:665-­‐677. Sudhof TC, Lottspeich F, Greengard P, Mehl E, Jahn R (1987) A synaptic vesicle protein with a novel cytoplasmic domain and four transmembrane regions. Science 238:1142-­‐1144. Sugita S, Janz R, Sudhof TC (1999) Synaptogyrins regulate Ca2+-­‐dependent exocytosis in PC12 cells. J Biol Chem 274:18893-­‐18901. Sullivan PF, Kendler KS, Neale MC (2003) Schizophrenia as a complex trait: evidence from a meta-­‐analysis of twin studies. Arch Gen Psychiatry 60:1187-­‐1192. Sutton RB, Fasshauer D, Jahn R, Brunger AT (1998) Crystal structure of a SNARE complex involved in synaptic exocytosis at 2.4 A resolution. Nature 395:347-­‐353. Sweitzer SM, Hinshaw JE (1998) Dynamin undergoes a GTP-­‐dependent conformational change causing vesiculation. Cell 93:1021-­‐1029. Takamori S et al. (2006) Molecular anatomy of a trafficking organelle. Cell 127:831-­‐846. Takei K, Haucke V (2001) Clathrin-­‐mediated endocytosis: membrane factors pull the trigger. Trends Cell Biol 11:385-­‐391. 62 Takei K, Mundigl O, Daniell L, De Camilli P (1996) The synaptic vesicle cycle: a single vesicle budding step involving clathrin and dynamin. J Cell Biol 133:1237-­‐1250. Tarpey PS et al. (2009) A systematic, large-­‐scale resequencing screen of X-­‐chromosome coding exons in mental retardation. Nat Genet 41:535-­‐543. Tebar F, Sorkina T, Sorkin A, Ericsson M, Kirchhausen T (1996) Eps15 is a component of clathrin-­‐coated pits and vesicles and is located at the rim of coated pits. J Biol Chem 271:28727-­‐28730. Teng H, Lin MY, Wilkinson RS (2007) Macroendocytosis and endosome processing in snake motor boutons. J Physiol 582:243-­‐262. Thiele C, Hannah MJ, Fahrenholz F, Huttner WB (2000) Cholesterol binds to synaptophysin and is required for biogenesis of synaptic vesicles. Nat Cell Biol 2:42-­‐49. Thomas L, Betz H (1990) Synaptophysin binds to physophilin, a putative synaptic plasma membrane protein. J Cell Biol 111:2041-­‐2052. Thomas L, Hartung K, Langosch D, Rehm H, Bamberg E, Franke WW, Betz H (1988) Identification of synaptophysin as a hexameric channel protein of the synaptic vesicle membrane. Science 242:1050-­‐1053. Tsukita S, Furuse M, Itoh M (2001) Multifunctional strands in tight junctions. Nat Rev Mol Cell Biol 2:285-­‐293. Ubach J, Zhang X, Shao X, Sudhof TC, Rizo J (1998) Ca2+ binding to synaptotagmin: how many Ca2+ ions bind to the tip of a C2-­‐domain? EMBO J 17:3921-­‐3930. Valtorta F, Meldolesi J, Fesce R (2001) Synaptic vesicles: is kissing a matter of competence? Trends Cell Biol 11:324-­‐328. Valtorta F, Pennuto M, Bonanomi D, Benfenati F (2004) Synaptophysin: leading actor or walk-­‐on role in synaptic vesicle exocytosis? Bioessays 26:445-­‐453. Verhage M, Maia AS, Plomp JJ, Brussaard AB, Heeroma JH, Vermeer H, Toonen RF, Hammer RE, van den Berg TK, Missler M, Geuze HJ, Sudhof TC (2000) Synaptic assembly of the brain in the absence of neurotransmitter secretion. Science 287:864-­‐869. Verma R, Kubendran S, Das SK, Jain S, Brahmachari SK (2005) SYNGR1 is associated with schizophrenia and bipolar disorder in southern India. J Hum Genet 50:635-­‐
640. Verma R, Chauhan C, Saleem Q, Gandhi C, Jain S, Brahmachari SK (2004) A nonsense mutation in the synaptogyrin 1 gene in a family with schizophrenia. Biol Psychiatry 55:196-­‐199. Verstreken P, Ohyama T, Bellen HJ (2008) FM 1-­‐43 labeling of synaptic vesicle pools at the Drosophila neuromuscular junction. Methods Mol Biol 440:349-­‐369. Verstreken P, Kjaerulff O, Lloyd TE, Atkinson R, Zhou Y, Meinertzhagen IA, Bellen HJ (2002) Endophilin mutations block clathrin-­‐mediated endocytosis but not neurotransmitter release. Cell 109:101-­‐112. Verstreken P, Koh TW, Schulze KL, Zhai RG, Hiesinger PR, Zhou Y, Mehta SQ, Cao Y, Roos J, Bellen HJ (2003) Synaptojanin is recruited by endophilin to promote synaptic vesicle uncoating. Neuron 40:733-­‐748. von Poser C, Zhang JZ, Mineo C, Ding W, Ying Y, Sudhof TC, Anderson RG (2000) Synaptotagmin regulation of coated pit assembly. J Biol Chem 275:30916-­‐30924. 63 Wang CT, Lu JC, Bai J, Chang PY, Martin TF, Chapman ER, Jackson MB (2003) Different domains of synaptotagmin control the choice between kiss-­‐and-­‐run and full fusion. Nature 424:943-­‐947. Washbourne P, Schiavo G, Montecucco C (1995) Vesicle-­‐associated membrane protein-­‐
2 (synaptobrevin-­‐2) forms a complex with synaptophysin. Biochem J 305 ( Pt 3):721-­‐724. Weber T, Zemelman BV, McNew JA, Westermann B, Gmachl M, Parlati F, Sollner TH, Rothman JE (1998) SNAREpins: minimal machinery for membrane fusion. Cell 92:759-­‐772. Whittaker VP, Michaelson IA, Kirkland RJ (1964) The separation of synaptic vesicles from nerve-­‐ending particles ('synaptosomes'). Biochem J 90:293-­‐303. Wiedenmann B, Franke WW (1985) Identification and localization of synaptophysin, an integral membrane glycoprotein of Mr 38,000 characteristic of presynaptic vesicles. Cell 41:1017-­‐1028. Windoffer R, Borchert-­‐Stuhltrager M, Haass NK, Thomas S, Hergt M, Bulitta CJ, Leube RE (1999) Tissue expression of the vesicle protein pantophysin. Cell Tissue Res 296:499-­‐510. Xue M, Stradomska A, Chen H, Brose N, Zhang W, Rosenmund C, Reim K (2008) Complexins facilitate neurotransmitter release at excitatory and inhibitory synapses in mammalian central nervous system. Proc Natl Acad Sci U S A 105:7875-­‐7880. Xue M, Lin YQ, Pan H, Reim K, Deng H, Bellen HJ, Rosenmund C (2009) Tilting the balance between facilitatory and inhibitory functions of mammalian and Drosophila Complexins orchestrates synaptic vesicle exocytosis. Neuron 64:367-­‐
380. Xue M, Reim K, Chen X, Chao HT, Deng H, Rizo J, Brose N, Rosenmund C (2007) Distinct domains of complexin I differentially regulate neurotransmitter release. Nat Struct Mol Biol 14:949-­‐958. Ye W, Lafer EM (1995) Bacterially expressed F1-­‐20/AP-­‐3 assembles clathrin into cages with a narrow size distribution: implications for the regulation of quantal size during neurotransmission. J Neurosci Res 41:15-­‐26. Young JD, Cohn ZA, Gilula NB (1987) Functional assembly of gap junction conductance in lipid bilayers: demonstration that the major 27 kd protein forms the junctional channel. Cell 48:733-­‐743. Zhang B, Koh YH, Beckstead RB, Budnik V, Ganetzky B, Bellen HJ (1998a) Synaptic vesicle size and number are regulated by a clathrin adaptor protein required for endocytosis. Neuron 21:1465-­‐1475. Zhang X, Rizo J, Sudhof TC (1998b) Mechanism of phospholipid binding by the C2A-­‐
domain of synaptotagmin I. Biochemistry 37:12395-­‐12403. Zhao H, Nonet ML (2001) A conserved mechanism of synaptogyrin localization. Mol Biol Cell 12:2275-­‐2289. Zhao W, Cavallaro S, Gusev P, Alkon DL (2000) Nonreceptor tyrosine protein kinase pp60c-­‐src in spatial learning: synapse-­‐specific changes in its gene expression, tyrosine phosphorylation, and protein-­‐protein interactions. Proc Natl Acad Sci U S A 97:8098-­‐8103. 64 Zhong Y, Wu CF (1991) Altered synaptic plasticity in Drosophila memory mutants with a defective cyclic AMP cascade. Science 251:198-­‐201. Zhou XW, Pfahnl A, Werner R, Hudder A, Llanes A, Luebke A, Dahl G (1997) Identification of a pore lining segment in gap junction hemichannels. Biophys J 72:1946-­‐1953. 65 66 Figure 1. Endocytosis at the presynaptic terminal. Three main endocytic routes for synaptic vesicle recycling have been described. Pink vesicles indicate the reserve pool, while blue vesicles signify the recycling pool. Clathrin-­‐mediated endocytosis (CME) usually occurs at a site somewhat removed from the active zone and involves the formation of a clathrin coat. During kiss-­‐and-­‐run endocytosis (KR), neurotransmitters are released through a fusion pore that subsequently closes to directly regenerate a synaptic vesicle. Both kiss-­‐and-­‐run endocytosis and clathrin-­‐mediated endocytosis are proposed to replenish the recycling pool. Bulk endocytosis (BE) occurs during intense neuronal stimulation and involves the invagination of large regions of the plasma membrane. Synaptic vesicles then bud from internalized endocytic cisternae and are preferentially targeted to the reserve pool. Image adapted from (Rizzoli and Betz, 2005). 67 Figure 2. Structure of synaptophysin and synaptogyrin. Synaptophysin and synapto-­‐
gyrin are tetraspanning integral membrane proteins with cytoplasmic N-­‐ and C-­‐termini. Lighter shading indicates the MARVEL transmembrane domains. Both proteins contain tyrosine phosphorylation sites (asterisks), although the exact number and positions of these sites are unknown. Synaptophysin has two disulfide bonds between its intravesicular loops (black lines) and is N-­‐glycosylated. Synaptogyrin contains two conserved cysteine residues within its first intravesicular loop, although it is unknown whether these residues form a disulfide bond in vivo. Figure based on (Fernandez-­‐
Chacon and Sudhof, 1999). 68 Chapter 2 Characterization of the Drosophila Synaptogyrin Homolog 1
Robin J. Stevens and J. Troy Littleton1 1 The Picower Institute for Learning and Memory, Department of Biology, Massachusetts Institute of Technology, Cambridge, MA 02139 Robin Stevens performed the majority of the work described in this chapter. Dina Volfson generated the UAS-­‐SCAMP-­‐RFP flies. 69 Introduction Sustained neurotransmission depends on the continual renewal of a pool of synaptic vesicles that are ready to respond to action potential propagation into the presynaptic nerve terminal (reviewed in (Sudhof, 2004)). Following exocytosis, synaptic vesicle proteins and lipids must be reclaimed via endocytosis and refilled with neurotransmitter. Synaptic vesicles must then traffic to the plasma membrane, where they dock at active zones and are primed for subsequent rounds of fusion. All steps of the synaptic vesicle cycle require the coordination of hundreds of proteins, many of which are uniquely targeted to synaptic vesicles. Much of our knowledge about neurotransmission and synaptic vesicle recycling has come from reductionist approaches that identify proteins enriched in the brain and then characterize their function by generating mutants, performing inhibition or overexpression studies, or conducting other in vivo or in vitro assays. Synaptophysin and synaptogyrin were among the first synaptic vesicle proteins to be discovered over twenty years ago (Jahn et al., 1985; Wiedenmann and Franke, 1985; Baumert et al., 1990; Stenius et al., 1995). Despite decades of research, the role of these proteins in the synaptic vesicle cycle remains elusive. Synaptogyrin and synaptophysin are evolutionarily related members of the MARVEL domain family, which defines a group of tetraspanning membrane proteins involved in processes such as vesicle trafficking and membrane apposition events (Hubner et al., 2002; Sanchez-­‐Pulido et al., 2002). These two proteins are abundantly expressed on synaptic vesicles, with synaptophysin alone constituting approximately 10% of synaptic vesicle protein by mass in rats (Takamori et al., 2006). Due to their 70 enrichment on synaptic vesicles, it is likely that these proteins play a functional or regulatory role in synaptic vesicle trafficking, release, or retrieval. Indeed, work in a variety of systems has implicated synaptogyrin and synaptophysin in synaptic vesicle exocytosis, endocytosis, and vesicle biogenesis, suggesting that these proteins may regulate multiple aspects of the synaptic vesicle cycle. Synaptophysin can oligomerize into homomultimers that may form ion channels or proteinaceous fusion pores (Rehm et al., 1986; Thomas et al., 1988; Gincel and Shoshan-­‐Barmatz, 2002; Arthur and Stowell, 2007). Synaptophysin also has been implicated in regulating synaptic vesicle exocytosis, perhaps via its interaction with synaptobrevin/VAMP2, which prevents the association of synaptobrevin with other members of the SNARE complex (Edelmann et al., 1995; Washbourne et al., 1995). Whether synaptophysin and synaptogyrin promote or inhibit synaptic vesicle release is unclear, as studies using Xenopus oocytes and cultured Xenopus motor neurons concluded that synaptophysin positively regulates exocytosis (Alder et al., 1992b; Alder et al., 1992a; Alder et al., 1995), while work using PC12 cells engineered to secrete human growth hormone suggested that synaptogyrin and synaptophysin inhibit fusion (Sugita et al., 1999). Additional studies suggest that synaptophysin and synaptogyrin family members (physins and gyrins, respectively) may play a role in shaping lipid membranes and promoting vesicle biogenesis. Synaptophysin binds cholesterol, potentially promoting the formation of high-­‐curvature membranes required for vesicle biogenesis (Thiele et al., 2000). Moreover, cellugyrin, a non-­‐neuronal paralog of synaptogyrin, can increase the formation of synaptic-­‐like microvesicles (SLMVs) in neuroendocrine cells (Belfort et 71 al., 2005). Synaptophysin may also participate in synaptic vesicle recycling through a calcium-­‐dependent interaction with dynamin, as the inhibition of this interaction results in a smaller synaptic vesicle pool following high-­‐frequency stimulation (Daly et al., 2000; Daly and Ziff, 2002). Similarly, a recent study in cultured hippocampal neurons revealed that the kinetics of endocytosis are slowed during periods of sustained stimulation in the absence of synaptophysin (Kwon and Chapman, 2011). Although synaptophysin and synaptogyrin are evolutionarily conserved and abundant proteins, their absence results in relatively mild phenotypes. Synaptophysin/ synaptogyrin double knockout mice are viable and fertile with normal basal neurotransmission, but they display defects in several forms of synaptic plasticity, including paired-­‐pulse facilitation (PPF) and long-­‐term potentiation (LTP) (McMahon et al., 1996; Janz et al., 1999). Similarly, the initial characterization of a C. elegans triple knockout of synaptophysin, synaptogyrin, and the tetraspanning membrane protein SCAMP also revealed no detectable morphological, electrophysiological, or behavioral phenotypes aside from a slight increase in the number of clathrin-­‐coated synaptic vesicles (Abraham et al., 2006). In-­‐depth analysis of the C. elegans synaptogyrin single mutant identified slightly altered motility and changes in sensitivity to drugs affecting GABAergic and cholinergic neurons (Abraham et al., 2011). The genetic analysis of synaptophysin and synaptogyrin in mammals, while potentially more relevant to human neurophysiology, is complicated by the presence of four physin and four gyrin paralogs. The synaptophysin paralog synaptoporin and the synaptogyrin paralog synaptogyrin 3 are enriched in the brain, and therefore may be able to substitute for the loss of either synaptophysin or synaptogyrin (Knaus et al., 72 1990; Belizaire et al., 2004). Indeed, a synaptophysin knockout mouse had altered synaptic vesicle density and vesicle size in photoreceptors in the outer plexiform layer of the retina, a region that lacks synaptoporin (McMahon et al., 1996; Spiwoks-­‐Becker et al., 2001). Similarly, synaptogyrin and synaptophysin are to some extent functionally redundant, as electrophysiological defects in PPF and LTP were only observed in the double knockout mice (however, a slight decrease in post-­‐tetanic potentiation was observed synaptogyrin-­‐deficient mice) (Janz et al., 1999). We therefore sought to investigate synaptogyrin function in Drosophila, which lacks a synaptophysin homolog and has only one synaptogyrin isoform. This approach allows us to simplify genetic analysis and avoid compensation by other gyrins and physins and simultaneously provides us with access to a well-­‐characterized, stereotypic synaptic structure – the Drosophila larval neuromuscular junction (NMJ). Our initial characterization of Drosophila synaptogyrin revealed that it localizes presynaptically at the larval NMJ and is abundantly expressed in the central nervous system of larvae and adult flies. We generated a synaptogyrin knockout and found that mutant animals are viable, fertile, and behaviorally normal, indicating that synaptogyrin is not an essential protein in Drosophila. 73 Results Evolutionary analysis of synaptogyrin and synaptophysin Synaptogyrin and synaptophysin have been established as evolutionarily conserved proteins (Hubner et al., 2002; Abraham et al., 2006). We were interested in further examining the evolutionary emergence of these proteins in relation to other synaptic vesicle components. While many synaptic vesicle proteins including SNAREs, V-­‐ATPase, and Rab3 are ancient eukaryotic proteins (Srivastava et al., 2010), key proteins involved in regulated calcium-­‐dependent exocytosis such as synaptotagmin 1 and complexin appeared more recently in evolution. Synaptotagmin and complexin homologs are absent in plants and fungi but both are present in the placozoan Trichoplax adhaerens – a simple multicellular organism that lacks a nervous system but has several organized cell layers (Srivastava et al., 2008; Barber et al., 2009). A protein BLAST search revealed that synaptogyrin and synaptophysin homologs are also found in T. adhaerens, indicating they also evolved before the emergence of defined synapses. Interestingly, we also identified a putative synaptogyrin homolog in the unicellular choanoflagellate Monosiga brevicollis with 33% identity to human synaptogyrin 1; however, we were unable to find a homolog in the yeast Saccharomyces cerevisiae (Table 1) (King et al., 2008). Our BLAST search did not reveal a M. brevicollis synaptophysin homolog, raising the possibility that all gyrin and physin family members are descended from an ancestral protein that was more similar to synaptogyrin than synaptophysin. Since choanoflagellates are thought to be the closest extant unicellular relative of metazoans (King et al., 2008), the presence of a synaptogyrin homolog in M. brevicollis but not in yeast suggests a unique role for 74 synaptogyrin in early metazoan evolution. Interestingly, the sea anemone Nematostella vectensis does not have a synaptogyrin homolog, but it does have at least one synaptophysin homolog, suggesting either that some metazoans may not require both proteins or that the functions of physins and gyrins are interchangeable (Putnam et al., 2007). We also searched for homologs of several other MARVEL domain-­‐containing proteins and found that some, including MAL (myelin and lymphocyte protein), MYADM (myeloid-­‐associated differentiation marker gene), and occludin (a component of tight junctions), appear to be vertebrate-­‐specific proteins. This is unsurprising given that vertebrate myelin and the myelin-­‐like sheaths found in some invertebrates appear to have arisen through convergent evolution (Hartline and Colman, 2007). Similarly, invertebrates do not have tight junctions, but rather have analogous structures known as septate junctions (Furuse and Tsukita, 2006). We also examined the CMTM family (CKLF-­‐like MARVEL transmembrane domain-­‐containing family, where CKLF stands for chemokine-­‐like factor), a novel group of proteins with a highly conserved member named CMTM4 that appears to be involved in regulating the cell cycle and cellular growth (Plate et al., 2010). Interestingly, CMTM4 homologs are found in Drosophila, C. elegans, N. vectensis, and potentially in T. adhaerens, although this putative homolog has only three predicted transmembrane domains. Finally, we investigated the evolutionary conservation of the secretory carrier-­‐associated membrane proteins (SCAMPs), which lack the MARVEL domain but have the same membrane topology (Hubbard et al., 2000). A SCAMP homolog was found in every organism we examined with the exception of yeast. Interestingly, SCAMP homologs are also found in the plant kingdom, which 75 suggests that they perform a more ubiquitous role in multicellular organisms (Fernandez-­‐Chacon and Sudhof, 2000). Protein sequence alignment of representative synaptogyrin homologs confirmed the previously reported observation that these proteins share the highest degree of homology within their transmembrane regions and that there is a substantial amount of variability in the composition and length of the N-­‐ and C-­‐termini (Figure 1) (Hubner et al., 2002). Unlike synaptogyrin homologs from other species, the putative M. brevicollis synaptogyrin lacks the two conserved cysteine residues located between the first two transmembrane domains that may form a disulfide bond within the first luminal loop (Figure 1). Interestingly, the M. brevicollis synaptogyrin homolog is almost entirely comprised of the MARVEL domain, with its short N-­‐ and C-­‐termini each predicted to be less than 15 amino acids in length. Therefore, it is likely that the MARVEL domain itself has an important cellular function and that the N-­‐ and C-­‐termini of gyrins and physins may have elaborated and adapted over the course of evolution to perform additional or more specialized tasks in other organisms. The Drosophila genome encodes a single synaptogyrin homolog (CG10808) that is 42% identical to human synaptogyrin 1 and shares the same predicted membrane topology as other gyrin family members. Unlike many other invertebrates, including C. elegans and the red flour beetle Tribolium castaneum, the Drosophila genome lacks a synaptophysin homolog. However, the genome does encode other MARVEL-­‐domain containing proteins, including a CMTM4 homolog (CG15211) and Singles Bar (CG13011), a protein involved in myoblast fusion (Estrada et al., 2007). The lack of a synaptophysin homolog in Drosophila is somewhat surprising given its high degree of 76 conservation across species. In C. elegans, however, synaptophysin expression is largely restricted to muscle cells in the pharynx and anal sphincter (Abraham et al., 2006), while synaptogyrin is expressed in most neurons (Nonet, 1999; Abraham et al., 2011), indicating that synaptogyrin is likely to be the predominant synaptic vesicle MARVEL protein in nematodes. Drosophila synaptogyrin is present in neurons and targeted to synaptic vesicles To our knowledge, synaptogyrin expression has not yet been characterized in Drosophila. Therefore, we examined the localization of endogenous synaptogyrin protein by generating antisera against the recombinant C-­‐terminus of Drosophila synaptogyrin. A Western blot of Drosophila adult head extracts reveals a single major band at an observed molecular weight of approximately 29 kDa (Figure 2A). To examine the subcellular distribution of synaptogyrin, we performed velocity gradient fractionation experiments using 10-­‐30% sucrose gradients. Synaptogyrin co-­‐migrated with other known synaptic vesicle proteins, including synaptotagmin 1, cysteine string protein (CSP), and synaptobrevin, suggesting that Drosophila synaptogyrin is enriched on synaptic vesicles (Figure 2B). As expected for a presumptive synaptic vesicle protein, immunohistochemistry of third instar Drosophila larvae results in abundant punctate synaptogyrin expression throughout the brain and ventral nerve cord (Figure 3). Similarly, immunohisto-­‐
chemistry of adult brains reveals profuse synaptogyrin protein expression in the glomeruli of the antennal lobes as well as in the optic lobes (Figure 4). Lower levels of punctate synaptogyrin immunofluorescence are found throughout the adult brain. Synaptogyrin also localizes to the neuromuscular junctions of third instar larvae where 77 it colocalizes with synaptotagmin 1 (Figure 5A) and partially overlaps with the active zone marker bruchpilot (nc82) (Figure 5B-­‐C). Synaptogyrin staining often is absent in the center of boutons, which reflects the tendency for synaptic vesicles to concentrate on the periphery of boutons as observed via electron microscopy (Atwood et al., 1993). Taken together, these results indicate that, like other synaptogyrin homologs, Drosophila synaptogyrin is a neuronal synaptic vesicle protein. Next, we utilized the GAL4/UAS system to compare the neuronal localization of GFP-­‐tagged Drosophila synaptogyrin (gyrin-­‐GFP) and RFP-­‐tagged Drosophila SCAMP (SCAMP-­‐RFP) driven by the pan-­‐neuronal driver elavc155-­GAL4 (Brand and Perrimon, 1993). SCAMP-­‐RFP and gyrin-­‐GFP extensively colocalize at the larval NMJ (Figure 6A), but show a more limited degree of overlap in the larval ventral nerve cord, with gyrin-­‐
GFP generally displaying a more restricted expression pattern (Figure 6B). In the large secretory cells of the salivary glands, SCAMP-­‐RFP and gyrin-­‐GFP also display a significant amount of colocalization (Figure 6C). However, gyrin-­‐GFP fluorescence is absent from many large SCAMP-­‐RFP-­‐positive puncta, suggesting that Drosophila SCAMP and synaptogyrin have different localization patterns. While these experiments do not indicate the endogenous expression pattern of Drosophila SCAMP, they do suggest that synaptogyrin and SCAMP are to some extent targeted to different subcellular compartments. Generation of a Drosophila synaptogyrin mutant We next sought to investigate the function of Drosophila synaptogyrin by creating and characterizing a synaptogyrin null mutant. We isolated two independent partial deletions of the synaptogyrin genomic locus via imprecise excision of a P-­‐
78 element insertion located approximately 500 bp upstream of the synaptogyrin translation start site. The first deletion, gyr1, extends 2.5 kb into the synaptogyrin locus and removes the first two exons and a portion of the third exon, while the second deletion, gyr2, is a smaller 1.7 kb deletion that removes the first exon (Figure 7A). A precise excision line, gyrPE, was chosen to serve as the control for genetic background in all experiments. The extent of the deletions and the precise excision event were confirmed by PCR and sequencing (Figure 7B). A Western blot of protein extracts revealed the complete absence of synaptogyrin immunoreactivity in both gyr1 and gyr2 animals (Figure 7C). Similarly, synaptogyrin antibody staining was absent at the neuromuscular junction in gyr third instar larvae (Figure 7D). The antibody raised against synaptogyrin targets the C-­‐terminus of the protein, and since the coding sequence of this region was left intact in both deletion lines, it is unlikely that a truncated version of the protein from a downstream start site is being produced. We therefore conclude that gyr1 and gyr2 are likely to be null mutations. Unless otherwise noted, gyr1 animals were used in all experiments. In agreement with nematode and mouse knockouts, Drosophila gyr animals are viable, fertile, and appear behaviorally normal. We examined the protein levels of a variety of synaptic proteins in adult head extracts and found that there were no dramatic alterations in protein expression (Figure 8A). Since synaptophysin influences synaptobrevin localization in cultured neurons (Pennuto et al., 2003; Gordon et al., 2011) and synaptogyrin promotes the targeting of synaptophysin to microvesicles in PC12 cells (Belfort and Kandror, 2003), we examined whether the loss of synaptogyrin affected the localization of synaptobrevin or synaptotagmin 1 in Drosophila. Both 79 synaptotagmin and synaptobrevin are correctly targeted to boutons at the larval NMJ in gyr mutants, with no apparent protein mislocalization to axons or inter-­‐bouton regions (Figure 8B). Therefore, we conclude that the loss of synaptogyrin does not dramatically affect the expression or localization of synaptotagmin and synaptobrevin. However, it remains possible that other synaptic proteins are mislocalized in gyr animals. We then asked whether the loss of synaptogyrin might affect the development or overall health of gyr mutants. When cultured in parallel, gyr mutants and controls proceed through the larval instar stages, pupate, and eclose at a similar rate, indicating no developmental delay (data not shown). We assayed the viability of gyr animals at 25oC by generating lifespan curves and quantifying the T50 (age at which 50% of the animals have died) for control (gyrPE) and gyr adult males and females, as well as for synaptogyrin rescues (elavc155-­GAL4; gyr; UAS-­myc-­gyrin/+) and gyr with the GAL4 driver alone (elavc155-­GAL4; gyr). The T50 values were nearly identical for control (males = 71 days, females = 63 days) and gyr (males = 71 days, females = 62 days), indicating that gyr does not cause a decrease in lifespan (Figure 8C). The presence of the elavc155-­
GAL4 driver slightly increased life expectancy, with both driver alone (males = 78 days, females = 67 days) and rescue animals (males = 79 days, females = 73 days) having a higher T50 than control and gyr animals. This is most likely due to differences in genetic backgrounds, as the elavc155-­GAL4 lines are in the white+ background, while control and gyr animals are white-­. Nevertheless, we conclude that synaptogyrin does not significantly impact the viability, fertility, or lifespan of Drosophila. 80 Behavioral analysis of gyr mutants reveals no significant deficits in climbing ability or courtship Although synaptogyrin-­‐deficient animals do not display any noticeable motor defects or a decrease in life expectancy, we wondered whether gyr mutants might have more subtle changes in locomotion or behavior. We therefore performed a climbing assay to determine whether gyr mutants were impaired in their basic motor function. Drosophila adults normally display a strong negative geotactic response (i.e., they climb upward) if disturbed (Erlenmeyer-­‐Kimling and Hirsch, 1961; Desroches et al., 2010). Climbing ability can be impaired by a wide variety of factors, including advanced age (Gargano et al., 2005), exposure to certain metals (Bonilla-­‐Ramirez et al., 2011), or overexpression of α-­‐synuclein in a Drosophila model of Parkinson’s disease (Feany and Bender, 2000). We compared the climbing ability of two-­‐day-­‐old and two-­‐week-­‐old gyr and control adult males by gently knocking the flies to the bottom of a vial and measuring the fraction of flies that had climbed 3 cm over 10 s intervals (Figure 9). Control and mutant animals performed similarly at both ages tested (p > 0.05, Student’s t-­‐test), indicating that there is no dramatic change in locomotor function up to two weeks of age in gyr mutants. Although we did not explicitly test climbing ability beyond two weeks of age, during the lifespan analysis we did not observe significant differences between control and gyr animals with respect to their ability to walk or climb, suggesting that the loss of synaptogyrin does not severely impact motor coordination in aged animals. To further analyze the behavior of gyr mutants we turned to the well-­‐
characterized Drosophila male courtship ritual. During courtship a male performs a 81 series of innate, stereotyped behaviors including tracking a female, tapping her with his forelimbs, extending and vibrating a wing to produce a species-­‐specific wing song, licking the female’s genitalia, and bending his abdomen to attempt copulation (Hall, 1994; Greenspan and Ferveur, 2000). These behaviors require the use of multiple sensory modalities as the male integrates visual, auditory, gustatory, and olfactory cues to successfully court a receptive conspecific female. When a naïve male is exposed to an unreceptive courtship target (e.g., a recently mated female or another male), he will learn to suppress courtship behavior upon subsequent encounters (Siegel and Hall, 1979; Vaias et al., 1993; Siwicki and Ladewski, 2003). This courtship suppression, also known as courtship conditioning, can last for hours to days depending on the training protocol and therefore provides a method of assaying short-­‐ and long-­‐term learning and memory in Drosophila (Siegel and Hall, 1979; McBride et al., 1999). We examined courtship conditioning in gyr mutant and control animals using a training protocol in which virgin males are paired with a previously mated female in a food vial for five hours (McBride et al., 1999). During this training period the males court on-­‐and-­‐off and therefore experience multiple rounds of rejection, which is thought to mimic the spaced training required to establish long-­‐term memory in several Drosophila learning paradigms (Tully et al., 1994; Keleman et al., 2007). One day following training, the courtship behavior of trained males is compared to age-­‐matched naïve males to determine the extent of courtship suppression (Figure 10A). Courtship levels are quantified using the courtship index (CI), which is defined as the fraction of time a male spends engaged in courtship behaviors over the course of ten minutes or until successful copulation. 82 Naïve gyr animals have a similar CI compared to naïve control males, and visually their overall courtship behavior appears normal (Figure 10B; control = 0.68 ± 0.05; gyr = 0.73 ± 0.05, p > 0.05, Student’s t-­‐test). A majority of mutant males successfully copulated during the ten-­‐minute observation period (10 of 17), indicating that gyr males are not considerably impaired in their ability to complete the courtship ritual (13 of 17 control males copulated). Furthermore, both gyr and control males that underwent training with a mated female displayed a similar amount of courtship suppression (indicated by a decrease in the CI) one day following training, signifying that gyr animals are also capable of learning (Figure 10B; control = 0.40 ± 0.08; gyr = 0.47 ± 0.08, p > 0.05, Student’s t-­‐test). We also quantified the average time to courtship initiation and the average time to copulation (for those copulations occurring in less than ten minutes) and found no significant differences between control and gyr males (Figure 10 C-­‐D; p > 0.05, Student’s t-­‐test). However, males of both genotypes that had undergone training with a previously mated female took a significantly longer time to initiate courtship with the tester female compared to naïve males (Figure 10 C; p < 0.05, two-­‐way ANOVA). This is unsurprising given that training induced an overall decrease in courtship behavior, therefore trained males would be expected to take longer to begin courting. Taken together, these results reveal no significant differences in courtship behavior or the ability of gyr mutants to form long-­‐lasting (one day) courtship memory. While we cannot rule out defects in other behavioral or learning paradigms, synaptogyrin does not appear to be required for locomotion or courtship, two essential Drosophila behaviors. 83 Discussion Drosophila synaptogyrin is a synaptic vesicle protein In mammals, synaptogyrin and synaptophysin are highly enriched on synaptic vesicles, with little to no expression outside of neuronal or neuroendocrine cells (Wiedenmann and Franke, 1985; Baumert et al., 1990). However, both proteins have ubiquitously expressed non-­‐neuronal homologs (cellugyrin and pantophysin), while the synaptophysin homolog mitsugumin 29 is only expressed in skeletal muscle, kidney, and the small intestine (Shimuta et al., 1998; Komazaki et al., 1999; Hubner et al., 2002). Protein BLAST searches confirmed that Drosophila synaptogyrin is similar in sequence to mammalian synaptogyrin 1, but this does not guarantee that Drosophila synaptogyrin is also a synaptic vesicle protein. Indeed, the expression pattern of the C. elegans synaptophysin homolog is restricted to certain muscle cells (Abraham et al., 2006), which implies that synaptophysin function is more limited in nematodes. Our work suggests that Drosophila synaptogyrin is a true neuronal homolog as it is broadly expressed throughout the nervous system, including at the neuromuscular junction where it colocalizes with the synaptic vesicle protein synaptotagmin 1. However, we did not investigate whether synaptogyrin is restricted to certain neuronal cell types or whether it is expressed pan-­‐neuronally. Only faint, nonspecific staining is observed in the muscle, suggesting that synaptogyrin is not expressed postsynaptically at the NMJ. Nevertheless, we cannot rule out that synaptogyrin expression is also postsynaptic in the central nervous system, or that it is present in non-­‐neuronal cell types. However, synaptogyrin mRNA expression levels as described by FlyAtlas (http://flyatlas.org) confirm that synaptogyrin is enriched 10-­‐26 times in the brain 84 relative to the whole fly, with little to no synaptogyrin mRNA present in non-­‐neuronal tissues (Chintapalli et al., 2007). Drosophila synaptogyrin comigrates with other known synaptic vesicle proteins in sucrose density gradients, suggesting that it localizes to synaptic vesicles. The extent to which synaptogyrin is targeted to other intracellular membranes is currently unknown, although we expect that some fraction of synaptogyrin is found on endosomes and at the plasma membrane as a result of synaptic vesicle exo-­‐endocytosis and normal intracellular trafficking. Neuronal coexpression of tagged synaptogyrin and SCAMP transgenes revealed a high amount of colocalization at the larval NMJ and ventral nerve cord, suggesting that Drosophila SCAMP, like several mammalian SCAMPs, is also targeted to synaptic vesicles (Fernandez-­‐Chacon and Sudhof, 2000; Hubner et al., 2002). However, SCAMP-­‐RFP also was found in many regions that lacked gyrin-­‐GFP, indicating that it may associate with a greater range of subcellular structures and therefore may play a more general role in intracellular trafficking. The endogenous expression pattern of Drosophila SCAMP is currently unknown, but its mRNA appears to be broadly expressed throughout the animal (http://flyatlas.org), suggesting that its function is not limited to neurons. The loss of synaptogyrin does not significantly impact certain adult behaviors The behavioral assays we undertook suggest that gyr adult flies are highly coordinated and that males are capable of effective courtship behavior. Although we cannot rule out subtle courtship defects such as alterations in wing song vibration, gyr males were capable of performing all aspects of the courtship ritual. The finding that gyr males are able to remember prior courtship encounters for at least one day suggests 85 that synaptogyrin is not required for all types of learning and memory in flies. However, it is currently unknown whether gyr mutants have defects in retaining memories for periods beyond one day. To examine this possibility, we attempted the same courtship conditioning assay with a four-­‐day interval between training and testing, but we only observed a slight, non-­‐significant drop in the CI of trained males compared to naïve males in both control and gyr animals. Changes to our experimental protocol, such as testing a larger number of males or using an immobilized female (which also strongly inhibits courtship), might allow us to determine whether the loss of synaptogyrin impacts long-­‐term memory beyond one day (Siegel and Hall, 1979). Similarly, other behavioral assays for testing long-­‐term memory could be employed, such as the classical conditioning paradigm in which olfactory cues are paired with a shock (Tully and Quinn, 1985; Tully et al., 1994). Work in mice has implicated synaptophysin in certain aspects of learning and memory including spatial learning and object novelty recognition (Schmitt et al., 2009). However, synaptogyrin’s impact on learning and memory in mice has not been investigated. Synaptogyrin is not required for viability, fertility, or basic motor function Our results indicate that synaptogyrin is not essential for viability, fertility, basic motor function, or courtship in Drosophila. This is consistent with knockout studies in other species that have also identified relatively mild phenotypes due to the loss of synaptogyrin and/or synaptophysin. Functional redundancy between synaptogyrin and synaptophysin has been documented in mice (Janz et al., 1999). Therefore, it is possible that the lack of severe phenotypes in flies, mice, or nematodes is due to compensation by some as yet unidentified protein or proteins. SCAMPs, although 86 lacking a MARVEL domain, have historically been linked with physins and gyrins because they share the same transmembrane topology and some SCAMP paralogs are located on synaptic vesicles (Brand et al., 1991; Fernandez-­‐Chacon and Sudhof, 2000; Hubner et al., 2002). Therefore, one or more SCAMPs potentially could compensate for the loss of gyrins and physins. However, the C. elegans synaptogyrin/synaptophysin/ SCAMP triple knockout demonstrates that the absence of all of these protein families does not result in a dramatic neuronal phenotype (Abraham et al., 2006). Since mice have four physins, four gyrins, and five SCAMPs, it is impractical to resolve this issue using the murine model. Even if one were to focus only on the proteins highly enriched in neurons, the total would still be six genes – synaptophysin, synaptoporin, synaptogyrins 1 and 3, and SCAMPs 1 and 5 (Grabs et al., 1994; Kedra et al., 1998; Fernandez-­‐Chacon and Sudhof, 2000; Belizaire et al., 2004). Drosophila has only a single SCAMP homolog and is therefore amenable to genetic analysis to examine issues of redundancy in future studies, although the work in C. elegans suggests it is unlikely that a Drosophila synaptogyrin/SCAMP double knockout would have severe neuronal impairments. 87 Methods Homology searches and protein alignment NCBI BLAST (blastp) was used to identify potential homologs, which were confirmed using a reciprocal blastp search (http://blast.ncbi.nlm.nih.gov). TMHMM Server v. 2.0 (http://www.cbs.dtu.dk/services/TMHMM) was used to verify that the prospective homologs had four transmembrane domains in the correct orientation (cytoplasmic N-­‐ and C-­‐termini). The synaptogyrin protein alignment was performed using the T-­‐Coffee algorithm (http://tcoffee.org) with default settings (Notredame et al., 2000; Di Tommaso et al., 2011). CLC DNA Workbench 4.0 was used to visualize the alignment. GenBank accession numbers used in the analysis are as follows: Synaptogyrin family: Drosophila melanogaster, AAF58329; Homo sapiens (synaptogyrin 1a), EAW60323; Mus musculus (synaptogyrin 1a), AAI38729; Danio rerio, AAI65917; Tribolium castaneum, EFA07832; Caenorhabditis elegans, AAC27798; Trichoplax adhaerens, EDV27716; Monosiga brevicollis, EDQ85723 Synaptophysin family: H. sapiens, AAB92358; M. musculus, EDL33904; D. rerio, CAQ13974; T. castaneum, EFA10995; C. elegans, AAB92070; N. vectensis, EDO44392 and EDO43817; T. adhaerens, EDV29003 CMTM (CKLF-­‐like MARVEL transmembrane domain-­‐containing) family: H. sapiens (CMTM4 isoform 1), AAN73845; M. musculus (CMTM4), AAN73437; D. rerio (CMTM4), AAH76315; D. melanogaster, AAF47955; T. castaneum, NCBI Reference Sequence XP_972350; C. elegans, AAB52340 (the exact protein length is disputed, see also AAV58868); N. vectensis, EDO45986; T. adhaerens, EDV25996 88 Occludin family: H. sapiens, AAC50451; M. musculus, AAC52515; D. rerio, AAH49304 MAL family: H. sapiens, AAA36196; M. musculus, AAH06826; D. rerio, AAH93153 MYADM family: H. sapiens, AAH13995; M. musculus, AAH56355; D. rerio, CAQ13223 SCAMP family: H. sapiens (SCAMP 1), AAH15065; M. musculus (SCAMP 1), AAH34283; D. rerio (SCAMP 1), AAH65684; D. melanogaster, AAF64492; T. castaneum, EFA11500; C. elegans, AAF36687; N. vectensis, EDO31962; T. adhaerens, EDV26892; M. brevicollis, EDQ84443 Fly stocks and transgenics Flies were cultured using standard media and techniques at room temperature (~22oC) unless otherwise noted. The GAL4/UAS system was utilized to drive neuronal expression of selected transgenes (Brand and Perrimon, 1993). The UAS-­‐synaptogyrin-­‐
GFP (gyrin-­‐GFP) and UAS-­‐SCAMP-­‐RFP (SCAMP-­‐RFP) constructs were subcloned into pUAST vectors using standard subcloning techniques and were injected into white (w1118) embryos. The UAS-­‐myc-­‐gyrin construct used for rescue experiments was subcloned into a modified pValum vector with an N-­‐terminal myc tag (Cho et al., 2010). This construct was injected into the strain yv;;attP2, which contains a site for targeted transgene insertion on the third chromosome (Markstein et al., 2008; Ni et al., 2008). The pan-­‐neuronal driver elavc155-­GAL4 was used to express the transgenes in neurons. Embryo injections were performed at Duke University Model Systems Genomics (Durham, NC) and Genetic Services, Inc. (Cambridge, MA). PCR primers were as follows: gyrin-­‐GFP 5’: 5’ – CGG AAT TCG GCG CCG GCG ATG AGT TCA C – 3’ gyrin-­‐GFP 3’: 5’ – CTA CTT ATG CGG CCG CAT AGG TGG GCT GCT GGT ACT – 3’ 89 myc-­‐gyrin 5’: 5’ – CGC ATA TGG ACA TGC TCA ACC AGA TAC TC – 3’ myc-­‐gyrin 3’: 5’ – GCT CTA GAC AAG TGT GGT AAT TCC TTA ATA G – 3’ Generation of a synaptogyrin null mutant Two independent partial deletions of the synaptogyrin locus were generated by imprecise excision of a P-­‐element (P{lacW}l(2)SH0644SH0644) located approximately 500 bp upstream of the synaptogyrin translation start site in the first exon of the gene. This P-­‐element insertion is listed as lethal on FlyBase (http://www.flybase.org). However, we were able to obtain viable and fertile homozygous animals, which suggests that either an unrelated lethal mutation was lost or a suppressor emerged. Approximately 200 white-­ excision events were screened by PCR, two of which resulted in deletions that extended into the synaptogyrin locus beyond the translation start site. The first, gyr1, is a 2.5 kb deletion that removes the first two exons and part of the third exon; the second, gyr2, is a 1.7 kb deletion that removes the first exon. A precise excision line isolated from the screen, gyrPE, was confirmed to be a precise excision via sequencing and was used as a control for genetic background in all experiments. To confirm that the deletions in gyr1 and gyr2 were indeed null mutations, we performed Western blots on protein extracts from these lines as well as from white (w1118) controls and gyrPE. The polyclonal antibody raised against the synaptogyrin protein targets the C-­‐terminus, and this coding region was left intact in both alleles. Synaptogyrin immunoreactivity was completely absent in gyr1 and gyr2 animals, indicating that a downstream start site is not producing a truncated version of the protein. gyr1 and gyr2 are in the white (w1118) background unless otherwise indicated. PCR primers used for the screen were as follows: 90 Across the P-­‐element insertion site: Primer 1: 5’ – GTC CAT GGT GAT GAT GGG TCT CTG ATG – 3’ Primer 2: 5’ – CAA TAT ATC TTG GGA GCT CTG CTG – 3’ Sequencing primers for deletions: Primer 1: 5’ – CGG AGG AGC AAG TGT GGT AAT TC – 3’ Primer 2: 5’ – CAA TAT ATC TTG GGA GCT CTG CTG – 3’ Synaptogyrin antibody generation The C-­‐terminal fragment of Drosophila synaptogyrin encoding amino acids 181-­‐
241 (gyrin-­‐cterm) was amplified from a Drosophila cDNA library and subcloned into the pGEX-­‐4T-­‐1 vector (GE Healthcare) using the EcoRI and NotI restriction sites. Recombinant GST-­‐gyrin-­‐cterm was expressed and processed in E. coli (BL21) using standard protocols. The fusion protein was purified from cell lysates using Glutathione Sepharose 4B (Amersham Biosciences) and was then used to immunize rabbits to generate polyclonal antibodies (Invitrogen). The synaptogyrin antibody serum recognizes a single major band at approximately 29 kDa on Western blots. PCR primers for cloning into the pGEX4T-­‐1 vector were as follows: gyrin-­‐cterm 5’: 5’ – CGG AAT TCG GCG CCG GCG ATG AGT TCA C – 3’ gyrin-­‐cterm 3’: 5’ – CTA GTT ATG CGG CCG CAT AGG TGG GCT GCT GGT ACT – 3’ Immunohistochemistry Wandering third instar larvae or pharate adults were dissected in calcium-­‐free HL3.1 saline (70 mM NaCl, 5 mM KCl, 10 mM NaHCO3, 4 mM MgCl2, 5 mM trehalose, 115 mM sucrose, 5 mM HEPES, pH 7.2) and fixed for 45 minutes in HL3.1 containing 4% 91 formaldehyde. Fixations using the synaptogyrin antibody (α-­‐Gyr) were fixed for 5 minutes in ice-­‐cold 100% methanol. Following several washes in phosphate-­‐buffered saline (PBS) and PBST (PBS with 0.1% Tween 20), larvae were incubated with primary antibodies in PBST overnight at 4oC. After additional washes, larvae were incubated in secondary antibodies in PBST for four hours at room temperature, washed, and mounted in 70% glycerol. The dilutions for primary antibodies were: synaptogyrin, 1:500; nc82/bruchpilot, 1:100; synaptotagmin 1 (monoclonal), 1:200 (3H2 2D7, from K. Zinn); synaptotagmin 1 (polyclonal), 1:500 (Littleton et al., 1993); and synaptobrevin, 1:500. The nc82 antibody developed by Erich Buchner was obtained from the Developmental Studies Hybridoma Bank developed under the auspices of the NICHD and maintained by the University of Iowa, Department of Biological Sciences, Iowa City, IA 52242. Secondary antibodies obtained from Jackson ImmunoResearch were used at a dilution of 1:250 and include Cy2-­‐conjugated goat anti-­‐rabbit, Cy3-­‐conjugated goat anti-­‐rabbit, and Rhodamine Red-­‐conjugated donkey anti-­‐mouse. Goat α-­‐HRP antibodies conjugated to either fluorescein isothiocyanate (FITC) or Rhodamine Red were added as indicated with secondary antibodies and were used at a concentration of 1:10,000 or 1:500, respectively. Images were acquired using confocal microscopy (Axoplan 2; Carl Zeiss MicroImaging, Inc.) using PASCAL software (Carl Zeiss) with 40x, 63x, or 100x oil-­‐
immersion lenses. Density gradient centrifugation The density gradient experiments were based on (van de Goor et al., 1995) and (Adolfsen et al., 2004). Buffer A (150 mM NaCl, 10 mM HEPES, 1 mM EGTA, 0.1 mM 92 MgCl2) was used as the lysis buffer and to make the sucrose gradients. Approximately 5 mL of adult flies of were collected on ice, flash frozen in liquid nitrogen, and vortexed vigorously to dislodge the heads from the bodies. Heads were collected using a pre-­‐
cooled sieve and homogenized on ice with a Dounce homogenizer in 1.5 mL buffer A with protease inhibitors (1 mM PMSF; 2 µg/mL aprotinin; 1 µg/mL leupeptin; 1 µg/mL pepstatin A; 1 mM EDTA). Post-­‐nuclear homogenates were then loaded onto 10-­‐30% sucrose gradients with a 50% sucrose pad and centrifuged in an SW 41 Ti rotor (Beckman Coulter) at 36,000 rpm for 4.5 hours at 4oC. Twenty 600 µL fractions were collected from the top of the gradient and were mixed with an equal volume of 2X SDS sample buffer prior to analysis by Western blot. Primary antibody concentrations were: synaptogyrin, 1:2,000; CSP, 1:500 (Zinsmaier et al., 1994); synaptotagmin 1 (DSyt2), 1:500 (Littleton et al., 1993); synaptobrevin, 1:1,000 (van de Goor et al., 1995); tubulin, 1:60,000; complexin, 1:2,000 (Huntwork and Littleton, 2007); HRS, 1:20,000 (Lloyd et al., 2002); syntaxin, 1:1,000 (8C3). The monoclonal antibody against syntaxin (8C3) developed by Seymour Benzer was obtained from the Developmental Studies Hybridoma Bank developed under the auspices of the NICHD and maintained by the University of Iowa, Department of Biological Sciences, Iowa City, IA 52242. The secondary antibodies IRDye 800-­‐conjugated goat anti-­‐mouse (LI-­‐COR Biosciences), IRDye 800-­‐conjugated goat anti-­‐rabbit (LI-­‐COR Biosciences), and IRDye 800-­‐conjugated goat anti-­‐guinea pig (Rockland Immunochemicals, Inc.) were used at a dilution of 1:5,000. Membranes were visualized using the LI-­‐COR Odyssey Imaging System (LI-­‐
COR Biosciences). 93 Western blots Western blots were performed using standard laboratory procedures. Nitrocellulose membranes were blocked for 30 minutes at room temperature in four parts PBS to one part Rockland Blocking Buffer (Rockland Immunochemicals, Inc.). Primary antibody incubation was done overnight at 4oC in a solution containing four parts PBST (PBS with 0.1% Tween 20) to one part Rockland Blocking Buffer. Primary antibody concentrations were: synaptogyrin, 1:20,000; synaptobrevin, 1:5,000 (van de Goor et al., 1995); SNAP 25, 1:1,000; Rab 3, 1:500; complexin, 1:5,000 (Huntwork and Littleton, 2007); CSP, 1:200 (Zinsmaier et al., 1994); syntaxin, 1:1,000 (8C3); synaptotagmin 1 (DSyt2), 1:10,000 (Littleton et al., 1993); Dlg, 1:1,000 (4F3); α-­‐SNAP, 1:1,000 (Babcock et al., 2004); ROP, 1:5,000 (4F8); arginine kinase, 1:10,000. The monoclonal antibodies against syntaxin (8C3), Dlg (4F3), and ROP (4F8) developed by Seymour Benzer, Corey Goodman, and Gerald Rubin, respectively, were obtained from the Developmental Studies Hybridoma Bank developed under the auspices of the NICHD and maintained by the University of Iowa, Department of Biological Sciences, Iowa City, IA 52242. The secondary antibodies IRDye 800-­‐conjugated goat anti-­‐mouse (LI-­‐COR Biosciences) and IRDye 800-­‐conjugated goat anti-­‐rabbit (LI-­‐COR Biosciences) were used at a dilution of 1:10,000. Visualization and quantification were performed using the LI-­‐COR Odyssey Imaging System (LI-­‐COR Biosciences). Lifespan analysis Approximately 200 males and 200 females of each genotype were collected shortly after eclosion, separated by sex, and grouped into 15 flies per vial. All flies were maintained at 25oC. The number of dead flies was recorded daily, and flies were 94 transferred to new vials three times per week. Flies lost or injured during transfer were removed from analysis. The totals at the end of the experiment were (female, male): gyrPE: 213, 196; gyr1: 211, 202; c155 control: 202, 182; rescue: 186, 190. c155 control genotype = elavc155-­GAL4/(Y or w); gyr1; rescue genotype = elavc155-­GAL4/(Y or w); gyr1; UAS-­myc-­gyrin/+ Climbing assay Males were collected shortly after eclosion and grouped with ten males per vial (four vials per genotype for a total of forty males per genotype) at 25oC. Flies were tested at two days and two weeks post-­‐eclosion and were transferred to new vials every 2-­‐3 days to avoid bacterial growth. On testing days, flies were transferred to empty vials with a line drawn 3 cm above the bottom of the vial. One control vial and one gyr vial were tested simultaneously. After allowing the flies to acclimatize for two minutes, the vials were gently banged 5-­‐6 times to knock the flies to the bottom of the vial, and the flies were given one minute to climb per trial. This process was repeated for a total of four trials for each pair of vials. Testing was videotaped, and the number of flies that had crossed the line was recorded for each trial. The four trials per vial were averaged, and then the vial averages were combined to give the genotype average. Courtship assays Courtship conditioning assays were based on (Keleman et al., 2007) and (McBride et al., 1999). All flies were raised, trained, and tested at 25oC. The gyr mutant and precise excision flies were generated in the white-­ background, which is known to alter courtship and learning (Campbell and Nash, 2001; Diegelmann et al., 2006). 95 Therefore, gyr strains were crossed into the white+ background for courtship assays. Virgin males were collected shortly after eclosion, isolated in single food vials, and aged 3-­‐4 days in a 12 hr:12 hr light-­‐dark cycle (both training and testing were performed during the light phase). Virgin Canton-­‐S females were aged 3-­‐5 days in groups of 10-­‐15 flies per vial. Canton-­‐S females used for training were mated with Canton-­‐S males (at least three days old) for ~20 hours the day before training and were removed approximately one hour prior to training. For the training phase, randomly chosen males were either paired with a pre-­‐
mated female in a food vial for five hours (trained), or placed in a food vial without a female for five hours (naïve). The cotton plug at the top of the food vial was pushed down to ~1 cm above the food surface to increase the frequency of contact between the male and female. Males were transferred to fresh food vials after training and were kept in isolation until testing. For the testing phase, individual males were paired with virgin Canton-­‐S females in a plastic mating chamber 1 cm in diameter for ten minutes. All tests were videotaped and manually scored for the courtship index, which is defined as the percentage of time a male engages in courtship behaviors during ten minutes or until copulation. Analysis was done blind to the genotype and experimental condition. 96 References Abraham C, Bai L, Leube RE (2011) Synaptogyrin-­‐dependent modulation of synaptic neurotransmission in Caenorhabditis elegans. Neuroscience 190:75-­‐88. Abraham C, Hutter H, Palfreyman MT, Spatkowski G, Weimer RM, Windoffer R, Jorgensen EM, Leube RE (2006) Synaptic tetraspan vesicle membrane proteins are conserved but not needed for synaptogenesis and neuronal function in Caenorhabditis elegans. Proc Natl Acad Sci U S A 103:8227-­‐8232. Adolfsen B, Saraswati S, Yoshihara M, Littleton JT (2004) Synaptotagmins are trafficked to distinct subcellular domains including the postsynaptic compartment. J Cell Biol 166:249-­‐260. Alder J, Xie ZP, Valtorta F, Greengard P, Poo M (1992a) Antibodies to synaptophysin interfere with transmitter secretion at neuromuscular synapses. Neuron 9:759-­‐
768. Alder J, Lu B, Valtorta F, Greengard P, Poo MM (1992b) Calcium-­‐dependent transmitter secretion reconstituted in Xenopus oocytes: requirement for synaptophysin. Science 257:657-­‐661. Alder J, Kanki H, Valtorta F, Greengard P, Poo MM (1995) Overexpression of synaptophysin enhances neurotransmitter secretion at Xenopus neuromuscular synapses. J Neurosci 15:511-­‐519. Arthur CP, Stowell MH (2007) Structure of synaptophysin: a hexameric MARVEL-­‐
domain channel protein. Structure 15:707-­‐714. Atwood HL, Govind CK, Wu CF (1993) Differential ultrastructure of synaptic terminals on ventral longitudinal abdominal muscles in Drosophila larvae. J Neurobiol 24:1008-­‐1024. Babcock M, Macleod GT, Leither J, Pallanck L (2004) Genetic analysis of soluble N-­‐
ethylmaleimide-­‐sensitive factor attachment protein function in Drosophila reveals positive and negative secretory roles. J Neurosci 24:3964-­‐3973. Barber CF, Jorquera RA, Melom JE, Littleton JT (2009) Postsynaptic regulation of synaptic plasticity by synaptotagmin 4 requires both C2 domains. J Cell Biol 187:295-­‐310. Baumert M, Takei K, Hartinger J, Burger PM, Fischer von Mollard G, Maycox PR, De Camilli P, Jahn R (1990) P29: a novel tyrosine-­‐phosphorylated membrane protein present in small clear vesicles of neurons and endocrine cells. J Cell Biol 110:1285-­‐1294. Belfort GM, Kandror KV (2003) Cellugyrin and synaptogyrin facilitate targeting of synaptophysin to a ubiquitous synaptic vesicle-­‐sized compartment in PC12 cells. J Biol Chem 278:47971-­‐47978. Belfort GM, Bakirtzi K, Kandror KV (2005) Cellugyrin induces biogenesis of synaptic-­‐
like microvesicles in PC12 cells. J Biol Chem 280:7262-­‐7272. Belizaire R, Komanduri C, Wooten K, Chen M, Thaller C, Janz R (2004) Characterization of synaptogyrin 3 as a new synaptic vesicle protein. J Comp Neurol 470:266-­‐281. Bonilla-­‐Ramirez L, Jimenez-­‐Del-­‐Rio M, Velez-­‐Pardo C (2011) Acute and chronic metal exposure impairs locomotion activity in Drosophila melanogaster: a model to study Parkinsonism. Biometals. 97 Brand AH, Perrimon N (1993) Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118:401-­‐415. Brand SH, Laurie SM, Mixon MB, Castle JD (1991) Secretory carrier membrane proteins 31-­‐35 define a common protein composition among secretory carrier membranes. J Biol Chem 266:18949-­‐18957. Campbell JL, Nash HA (2001) Volatile general anesthetics reveal a neurobiological role for the white and brown genes of Drosophila melanogaster. J Neurobiol 49:339-­‐
349. Chintapalli VR, Wang J, Dow JA (2007) Using FlyAtlas to identify better Drosophila melanogaster models of human disease. Nat Genet 39:715-­‐720. Cho RW, Song Y, Littleton JT (2010) Comparative analysis of Drosophila and mammalian complexins as fusion clamps and facilitators of neurotransmitter release. Mol Cell Neurosci 45:389-­‐397. Daly C, Ziff EB (2002) Ca2+-­‐dependent formation of a dynamin-­‐synaptophysin complex: potential role in synaptic vesicle endocytosis. J Biol Chem 277:9010-­‐9015. Daly C, Sugimori M, Moreira JE, Ziff EB, Llinas R (2000) Synaptophysin regulates clathrin-­‐independent endocytosis of synaptic vesicles. Proc Natl Acad Sci U S A 97:6120-­‐6125. Desroches CE, Busto M, Riedl CA, Mackay TF, Sokolowski MB (2010) Quantitative trait locus mapping of gravitaxis behaviour in Drosophila melanogaster. Genet Res (Camb) 92:167-­‐174. Di Tommaso P, Moretti S, Xenarios I, Orobitg M, Montanyola A, Chang JM, Taly JF, Notredame C (2011) T-­‐Coffee: a web server for the multiple sequence alignment of protein and RNA sequences using structural information and homology extension. Nucleic Acids Res 39:W13-­‐17. Diegelmann S, Zars M, Zars T (2006) Genetic dissociation of acquisition and memory strength in the heat-­‐box spatial learning paradigm in Drosophila. Learn Mem 13:72-­‐83. Edelmann L, Hanson PI, Chapman ER, Jahn R (1995) Synaptobrevin binding to synaptophysin: a potential mechanism for controlling the exocytotic fusion machine. EMBO J 14:224-­‐231. Erlenmeyer-­‐Kimling L, Hirsch J (1961) Measurement of the relations between chromosomes and behavior. Science 134:1068-­‐1069. Estrada B, Maeland AD, Gisselbrecht SS, Bloor JW, Brown NH, Michelson AM (2007) The MARVEL domain protein, Singles Bar, is required for progression past the pre-­‐
fusion complex stage of myoblast fusion. Dev Biol 307:328-­‐339. Feany MB, Bender WW (2000) A Drosophila model of Parkinson's disease. Nature 404:394-­‐398. Fernandez-­‐Chacon R, Sudhof TC (2000) Novel SCAMPs lacking NPF repeats: ubiquitous and synaptic vesicle-­‐specific forms implicate SCAMPs in multiple membrane-­‐
trafficking functions. J Neurosci 20:7941-­‐7950. Furuse M, Tsukita S (2006) Claudins in occluding junctions of humans and flies. Trends Cell Biol 16:181-­‐188. Gargano JW, Martin I, Bhandari P, Grotewiel MS (2005) Rapid iterative negative geotaxis (RING): a new method for assessing age-­‐related locomotor decline in Drosophila. Exp Gerontol 40:386-­‐395. 98 Gincel D, Shoshan-­‐Barmatz V (2002) The synaptic vesicle protein synaptophysin: purification and characterization of its channel activity. Biophys J 83:3223-­‐3229. Gordon SL, Leube RE, Cousin MA (2011) Synaptophysin Is Required for Synaptobrevin Retrieval during Synaptic Vesicle Endocytosis. J Neurosci 31:14032-­‐14036. Grabs D, Bergmann M, Schuster T, Fox PA, Brich M, Gratz M (1994) Differential expression of synaptophysin and synaptoporin during pre-­‐ and postnatal development of the rat hippocampal network. Eur J Neurosci 6:1765-­‐1771. Greenspan RJ, Ferveur JF (2000) Courtship in Drosophila. Annu Rev Genet 34:205-­‐232. Hall JC (1994) The mating of a fly. Science 264:1702-­‐1714. Hartline DK, Colman DR (2007) Rapid conduction and the evolution of giant axons and myelinated fibers. Curr Biol 17:R29-­‐35. Hubbard C, Singleton D, Rauch M, Jayasinghe S, Cafiso D, Castle D (2000) The secretory carrier membrane protein family: structure and membrane topology. Mol Biol Cell 11:2933-­‐2947. Hubner K, Windoffer R, Hutter H, Leube RE (2002) Tetraspan vesicle membrane proteins: synthesis, subcellular localization, and functional properties. Int Rev Cytol 214:103-­‐159. Huntwork S, Littleton JT (2007) A complexin fusion clamp regulates spontaneous neurotransmitter release and synaptic growth. Nat Neurosci 10:1235-­‐1237. Jahn R, Schiebler W, Ouimet C, Greengard P (1985) A 38,000-­‐dalton membrane protein (p38) present in synaptic vesicles. Proc Natl Acad Sci U S A 82:4137-­‐4141. Janz R, Sudhof TC, Hammer RE, Unni V, Siegelbaum SA, Bolshakov VY (1999) Essential roles in synaptic plasticity for synaptogyrin I and synaptophysin I. Neuron 24:687-­‐700. Kedra D, Pan HQ, Seroussi E, Fransson I, Guilbaud C, Collins JE, Dunham I, Blennow E, Roe BA, Piehl F, Dumanski JP (1998) Characterization of the human synaptogyrin gene family. Hum Genet 103:131-­‐141. Keleman K, Kruttner S, Alenius M, Dickson BJ (2007) Function of the Drosophila CPEB protein Orb2 in long-­‐term courtship memory. Nat Neurosci 10:1587-­‐1593. King N et al. (2008) The genome of the choanoflagellate Monosiga brevicollis and the origin of metazoans. Nature 451:783-­‐788. Knaus P, Marqueze-­‐Pouey B, Scherer H, Betz H (1990) Synaptoporin, a novel putative channel protein of synaptic vesicles. Neuron 5:453-­‐462. Komazaki S, Nishi M, Kangawa K, Takeshima H (1999) Immunolocalization of mitsugumin29 in developing skeletal muscle and effects of the protein expressed in amphibian embryonic cells. Dev Dyn 215:87-­‐95. Kwon SE, Chapman ER (2011) Synaptophysin regulates the kinetics of synaptic vesicle endocytosis in central neurons. Neuron 70:847-­‐854. Littleton JT, Bellen HJ, Perin MS (1993) Expression of synaptotagmin in Drosophila reveals transport and localization of synaptic vesicles to the synapse. Development 118:1077-­‐1088. Lloyd TE, Atkinson R, Wu MN, Zhou Y, Pennetta G, Bellen HJ (2002) Hrs regulates endosome membrane invagination and tyrosine kinase receptor signaling in Drosophila. Cell 108:261-­‐269. 99 Markstein M, Pitsouli C, Villalta C, Celniker SE, Perrimon N (2008) Exploiting position effects and the gypsy retrovirus insulator to engineer precisely expressed transgenes. Nat Genet 40:476-­‐483. McBride SM, Giuliani G, Choi C, Krause P, Correale D, Watson K, Baker G, Siwicki KK (1999) Mushroom body ablation impairs short-­‐term memory and long-­‐term memory of courtship conditioning in Drosophila melanogaster. Neuron 24:967-­‐
977. McMahon HT, Bolshakov VY, Janz R, Hammer RE, Siegelbaum SA, Sudhof TC (1996) Synaptophysin, a major synaptic vesicle protein, is not essential for neurotransmitter release. Proc Natl Acad Sci U S A 93:4760-­‐4764. Ni JQ, Markstein M, Binari R, Pfeiffer B, Liu LP, Villalta C, Booker M, Perkins L, Perrimon N (2008) Vector and parameters for targeted transgenic RNA interference in Drosophila melanogaster. Nat Methods 5:49-­‐51. Nonet ML (1999) Visualization of synaptic specializations in live C. elegans with synaptic vesicle protein-­‐GFP fusions. J Neurosci Methods 89:33-­‐40. Notredame C, Higgins DG, Heringa J (2000) T-­‐Coffee: A novel method for fast and accurate multiple sequence alignment. J Mol Biol 302:205-­‐217. Pennuto M, Bonanomi D, Benfenati F, Valtorta F (2003) Synaptophysin I controls the targeting of VAMP2/synaptobrevin II to synaptic vesicles. Mol Biol Cell 14:4909-­‐
4919. Plate M, Li T, Wang Y, Mo X, Zhang Y, Ma D, Han W (2010) Identification and characterization of CMTM4, a novel gene with inhibitory effects on HeLa cell growth through Inducing G2/M phase accumulation. Mol Cells 29:355-­‐361. Putnam NH, Srivastava M, Hellsten U, Dirks B, Chapman J, Salamov A, Terry A, Shapiro H, Lindquist E, Kapitonov VV, Jurka J, Genikhovich G, Grigoriev IV, Lucas SM, Steele RE, Finnerty JR, Technau U, Martindale MQ, Rokhsar DS (2007) Sea anemone genome reveals ancestral eumetazoan gene repertoire and genomic organization. Science 317:86-­‐94. Rehm H, Wiedenmann B, Betz H (1986) Molecular characterization of synaptophysin, a major calcium-­‐binding protein of the synaptic vesicle membrane. EMBO J 5:535-­‐
541. Sanchez-­‐Pulido L, Martin-­‐Belmonte F, Valencia A, Alonso MA (2002) MARVEL: a conserved domain involved in membrane apposition events. Trends Biochem Sci 27:599-­‐601. Schmitt U, Tanimoto N, Seeliger M, Schaeffel F, Leube RE (2009) Detection of behavioral alterations and learning deficits in mice lacking synaptophysin. Neuroscience 162:234-­‐243. Shimuta M, Komazaki S, Nishi M, Iino M, Nakagawara K, Takeshima H (1998) Structure and expression of mitsugumin29 gene. FEBS Lett 431:263-­‐267. Siegel RW, Hall JC (1979) Conditioned responses in courtship behavior of normal and mutant Drosophila. Proc Natl Acad Sci U S A 76:3430-­‐3434. Siwicki KK, Ladewski L (2003) Associative learning and memory in Drosophila: beyond olfactory conditioning. Behav Processes 64:225-­‐238. Spiwoks-­‐Becker I, Vollrath L, Seeliger MW, Jaissle G, Eshkind LG, Leube RE (2001) Synaptic vesicle alterations in rod photoreceptors of synaptophysin-­‐deficient mice. Neuroscience 107:127-­‐142. 100 Srivastava M et al. (2008) The Trichoplax genome and the nature of placozoans. Nature 454:955-­‐960. Srivastava M et al. (2010) The Amphimedon queenslandica genome and the evolution of animal complexity. Nature 466:720-­‐726. Stenius K, Janz R, Sudhof TC, Jahn R (1995) Structure of synaptogyrin (p29) defines novel synaptic vesicle protein. J Cell Biol 131:1801-­‐1809. Sudhof TC (2004) The synaptic vesicle cycle. Annu Rev Neurosci 27:509-­‐547. Sugita S, Janz R, Sudhof TC (1999) Synaptogyrins regulate Ca2+-­‐dependent exocytosis in PC12 cells. J Biol Chem 274:18893-­‐18901. Takamori S et al. (2006) Molecular anatomy of a trafficking organelle. Cell 127:831-­‐846. Thiele C, Hannah MJ, Fahrenholz F, Huttner WB (2000) Cholesterol binds to synaptophysin and is required for biogenesis of synaptic vesicles. Nat Cell Biol 2:42-­‐49. Thomas L, Hartung K, Langosch D, Rehm H, Bamberg E, Franke WW, Betz H (1988) Identification of synaptophysin as a hexameric channel protein of the synaptic vesicle membrane. Science 242:1050-­‐1053. Tully T, Quinn WG (1985) Classical conditioning and retention in normal and mutant Drosophila melanogaster. J Comp Physiol A 157:263-­‐277. Tully T, Preat T, Boynton SC, Del Vecchio M (1994) Genetic dissection of consolidated memory in Drosophila. Cell 79:35-­‐47. Vaias LJ, Napolitano LM, Tompkins L (1993) Identification of stimuli that mediate experience-­‐dependent modification of homosexual courtship in Drosophila melanogaster. Behav Genet 23:91-­‐97. van de Goor J, Ramaswami M, Kelly R (1995) Redistribution of synaptic vesicles and their proteins in temperature-­‐sensitive shibire(ts1) mutant Drosophila. Proc Natl Acad Sci U S A 92:5739-­‐5743. Washbourne P, Schiavo G, Montecucco C (1995) Vesicle-­‐associated membrane protein-­‐
2 (synaptobrevin-­‐2) forms a complex with synaptophysin. Biochem J 305 ( Pt 3):721-­‐724. Wiedenmann B, Franke WW (1985) Identification and localization of synaptophysin, an integral membrane glycoprotein of Mr 38,000 characteristic of presynaptic vesicles. Cell 41:1017-­‐1028. Zinsmaier KE, Eberle KK, Buchner E, Walter N, Benzer S (1994) Paralysis and early death in cysteine string protein mutants of Drosophila. Science 263:977-­‐980. 101 102 103 Figure 1. Protein sequence alignment of synaptogyrin homologs. An alignment of synaptogyrin homologs was generated using the T-­‐coffee algorithm (http://tcoffee.org) with default parameters (Notredame et al., 2000). The T-­‐coffee algorithm assigns a score based on the quality of the alignment, with red regions indicating a good alignment and blue signifying a region with poor alignment. Transmembrane (TM) domains are indicated for Drosophila synaptogyrin as determined by the trans-­‐
membrane region predictor TMHMM Server 2.0 (http://www.cbs.dtu.dk/services/ TMHMM). Arrows designate the two conserved cysteine residues between the first two transmembrane helices. Sequence abbreviations: D.m., Drosophila melanogaster; H.s., Homo sapiens; M.m., Mus musculus; T.c., Tribolium castaneum; C.e., Caenorhabditis elegans; T.a., Trichoplax adhaerens; M.b., Monosiga brevicollis. GenBank accession numbers are listed in Methods. 104 Figure 2. Subcellular localization of Drosophila synaptogyrin. A) A polyclonal synaptogyrin antibody recognizes a single major band at approximately 29 kDa in adult head homogenates. B) Post-­‐nuclear head homogenates were separated using a 10-­‐30% sucrose density gradient. Fractions were analyzed by Western blot using antibodies for various subcellular markers. Synaptotagmin 1 (syt1), synaptobrevin (n-­‐syb) and cysteine string protein (CSP) were used as synaptic vesicle fraction markers. Antisera against HRS were used to identify fractions containing endosomes (left-­‐most fractions), while syntaxin (syx) antibodies were used to mark the plasma membrane (right-­‐most fractions). Fractions were also immunostained with complexin (cpx) and tubulin (tub) antisera. Synaptogyrin (gyr) primarily co-­‐migrates with other synaptic vesicle proteins. 105 106 Figure 3. Synaptogyrin protein expression in larvae. A) Immunohistochemistry with antisera against the neuronal membrane marker horseradish peroxidase (HRP, green) and synaptogyrin (Gyr, red) indicates that synaptogyrin is broadly expressed throughout the central nervous system in third instar larvae. B) A higher magnification image of (A) reveals punctate synaptogyrin staining along the ventral nerve cord. C) One lobe of the larval brain from (A) at higher magnification. D) Synaptogyrin-­‐positive puncta are also present in motorneuron axons and likely mark synaptic vesicles trafficking to the neuromuscular junction. Scale bars: A = 100 μm; B and C = 50 μm; D = 20 μm. 107 108 Figure 4. Synaptogyrin protein expression in the adult brain. A) A schematic of the adult brain (anterior view) with the regions imaged in (B) and (C) indicated by boxes. The antennal lobes (al) and optic lobes (ol) are highlighted since these regions have high levels of synaptogyrin protein expression. B) Synaptogyrin immunostaining (Gyr, red) is located throughout the central brain regions with particularly high expression found in the glomeruli of the antennal lobes. Horseradish peroxidase staining (HRP, green) labels neuronal membranes. C) Synaptogyrin protein expression is abundant in the optic lobes of adult flies. Scale bars = 50 μm. 109 Figure 5. Synaptogyrin protein localization at larval NMJs. A) Synaptogyrin (Gyr, green) is expressed presynaptically at the larval NMJ where it colocalizes with synaptotagmin 1 (Syt1, magenta). B) Synaptogyrin is not confined to active zones, which are indicated by nc82 staining (magenta). C) A magnified view of the region indicated in (B) showing synaptogyrin’s localization within boutons. Scale bars: A = 25 μm; B = 20 μm; C = 2 μm. 110 111 Figure 6. Gyrin-­‐GFP and SCAMP-­‐RFP partially colocalize in the nervous system. A) SCAMP-­‐RFP (magenta) and gyrin-­‐GFP (green) almost completely colocalize at the third instar larval NMJ when driven with a pan-­‐neuronal GAL4 driver, suggesting that SCAMP is also targeted to synaptic vesicles. Both transgenes are enriched in boutons and are largely absent from the regions between boutons. B) SCAMP-­‐RFP and gyrin-­‐GFP have punctate staining along the larval ventral nerve cord (VNC), although gyrin-­‐GFP localization is more restricted, especially in the lateral regions of the VNC. C) Cells in the salivary gland (SG) also express the SCAMP and gyrin transgenes. While there is some degree of overlap, SCAMP-­‐RFP and gyrin-­‐GFP also appear to be trafficked to different subcellular compartments. Note that gyrin-­‐GFP is excluded from many large SCAMP-­‐RFP-­‐positive puncta (see inset). Scale bars: A = 20 μm; B and C = 50 μm. 112 113 Figure 7. Generation of a Drosophila synaptogyrin (gyr) mutant. A) The synaptogyrin locus is diagrammed with the two neighboring genes (CG30484 and CG6357). The location of the P-­‐element used for the excision screen is indicated by the pink triangle. Two separate deletions were isolated and are indicated by black lines. The green arrows mark the locations for the PCR primers used to determine the extent of each deletion. B) PCR products amplifying the region between the green arrows in (A) from genomic DNA isolated from white control, gyrPE, gyr1, and gyr2 adult flies. This region is approximately 3.5 kb in wild-­‐type flies (lacking the P-­‐element), while the gyr1 and gyr2 PCR products are ~1 kb and ~1.8 kb, respectively. Therefore, the gyr1 excision event is a ~2.5 kb deletion, while gyr2 is a ~1.7 kb deletion. PCR products were sequenced to confirm that gyrPE is indeed a precise excision and to verify that the deletions do not extend outside of the synaptogyrin locus. C) Synaptogyrin protein expression levels from homogenates made from white, gyrPE, gyr1, and gyr2 adult heads were determined by Western blot analysis. Immunostaining for complexin (cpx) was used as a loading control. Synaptogyrin immunoreactivity is absent in gyr1 and gyr2, and since the synaptogyrin polyclonal antibody was raised against a region of the protein downstream of both deletions, it is unlikely that a truncated version of synaptogyrin is being produced. We therefore conclude that gyr1 and gyr2 are most likely null mutations. D) Immunohistochemistry at third instar larval NMJs confirms that synaptogyrin (red) is absent in the gyr1 mutant but not in the control (gyrPE). Synaptic varicosities were identified using α-­‐HRP antibodies (green). 114 115 Figure 8. Examination of protein expression levels, synaptic vesicle protein localization, and viability in gyr mutants. A) The expression levels of a variety of synaptic proteins are unchanged in gyr head extracts relative to controls. Proteins analyzed include synaptobrevin (n-­‐syb), SNAP 25, Rab 3, complexin (Cpx), cysteine string protein (CSP), syntaxin (Syx), synaptotagmin 1 (Syt 1), Discs large (Dlg), soluble N-­‐ethylmaleimide-­‐sensitive fusion protein (NSF)-­‐attachment protein (α-­‐SNAP), sec1/unc-­‐18 (ROP), and arginine kinase (AK). B) The localization of synaptobrevin (upper panels) and synaptotagmin 1 (lower panels) is similar in gyr mutants compared to controls. Immunostaining of both proteins (green) is almost entirely restricted to synaptic varicosities with little trafficking to regions between boutons or along the axon. Axonal membranes are visualized with α-­‐HRP antibodies (red). C) Both gyr mutant males (left) and females (right) have a similar lifespan compared to controls (gyrPE). c155 control (elavc155-­GAL4; gyr) flies have a slightly increased lifespan, and female rescue flies (elavc155-­GAL4/w; gyr; UAS-­myc-­gyrin/+) live somewhat longer than c155 control females. The T50 (age at which 50% of the flies have died) in days for each sex and genotype were: males: control = 71, gyr = 71, c155 control = 78, rescue = 79; females: control = 63, gyr = 62, c155 control = 67, rescue = 73. 116 Figure 9. Locomotor ability as measured by a climbing assay is not significantly different in gyr mutants. A) Images captured from a video of a climbing assay trial. The vial on the left contains gyr mutant flies, while the vial on the right contains control flies. The numbers on the bottom left of the panels indicate the time in seconds after the flies were gently knocked to the bottom of the vials. After ten seconds most flies have crossed the black line drawn 3 cm above the bottom of the vial. B) Quantification of climbing assays. A total of forty males were analyzed in groups of ten, with four trials for each group. The percentage of flies that had crossed the 3 cm line at each time point (up to one minute in ten-­‐second intervals) is not significantly different between gyr mutants and controls at any point (p > 0.05, Student’s t-­‐test). Similar results were seen using flies aged two days (left) and aged two weeks (right). 117 118 Figure 10. Analysis of courtship behavior in gyr mutant males. A) Schematic of the assay used to assess courtship conditioning. Males were kept in isolation after eclosion and between training and testing. During the training phase, half the males were paired with a previously mated female and the other half were sham trained for five hours. The following day males were paired with a new female for ten minutes and the courtship index (CI) was determined. B) CIs for gyr and control males under naïve and trained conditions. There is no statistically significant difference in performance between gyr and control males under either training condition (p > 0.05, Student’s t-­‐
test). However, trained males of both genotypes display reduced courtship relative to naïve males (control, p = 0.005; gyr, p = 0.009, Student’s t-­‐test). Average CIs ± SEM: control, naïve = 0.68 ± 0.050, n = 17; control, trained = 0.40 ± 0.078, n = 17; gyr, naïve = 0.73 ± 0.051, n = 17; gyr, trained = 0.47 ± 0.078, n = 15. C) Analysis by two-­‐way ANOVA reveals that trained males of both genotypes have a delay in courtship initiation relative to naïve males (p < 0.05). However, there is no significant difference in time to courtship initiation between control and mutant flies under either training condition (p > 0.05, Student’s t-­‐test). Average time to initiation (in seconds) ± SEM: control, naïve = 48.4 ± 11.6, n = 17; control, trained = 104.6 ± 34.9, n = 17; gyr, naïve = 31.5 ± 8.01, n = 17; gyr, trained = 68.1 ± 21.8, n = 15. D) For the males that successfully copulated during the ten-­‐minute observation period there was no statistically significant difference in time to copulation (from the introduction of the female) between control and gyr males under both naïve and trained conditions (p > 0.05, Student’s t-­‐test). Average time to copulation (in seconds) ± SEM: control, naïve = 193.3 ± 31.3, n = 13; control, trained = 163.7 ± 40.7, n = 6; gyr, naïve = 255.3 ± 28.5, n = 10; gyr, trained = 222.4 ± 48.4, n = 8. 119 120 Chapter 3 Analysis of Synaptogyrin’s Role in Synaptic Vesicle Exo-­endocytosis 1
Robin J. Stevens , Yulia Akbergenova1, Ramon Jorquera1, and J. Troy Littleton1 1 The Picower Institute for Learning and Memory, Department of Biology, Massachusetts Institute of Technology, Cambridge, MA 02139 Robin Stevens performed the majority of the work described in this chapter. Yulia Akbergenova performed the electron microscopy. Ramon Jorquera performed the voltage clamp electrophysiology. 121 Introduction The lack of severe neuronal phenotypes in the mouse, C. elegans, and Drosophila synaptophysin and synaptogyrin knockouts is somewhat surprising given the degree of evolutionary conservation of these proteins as well as their relatively high expression levels on synaptic vesicles (Hubner et al., 2002; Takamori et al., 2006; Mutch et al., 2011). In mice, there is clearly some degree of redundancy between synaptophysin and synaptogyrin (Janz et al., 1999), and potentially between synaptophysin and synaptoporin as well (Spiwoks-­‐Becker et al., 2001). Furthermore, synaptogyrin 3 may compensate for the loss of these proteins, as it is highly enriched in the brain and targeted to synaptic vesicles (Belizaire et al., 2004). However, the absence of all physin and gyrin isoforms in C. elegans does not dramatically affect basal synaptic function or animal behavior (Abraham et al., 2006). Nevertheless, there is a growing body of evidence suggesting that these proteins, while not strictly required for exocytosis or endocytosis, are able to modulate certain aspects of synaptic vesicle release and recycling. In particular, it appears as though synaptophysin and/or synaptogyrin may participate in a clathrin-­‐independent version of synaptic vesicle endocytosis. An increase in clathrin-­‐coated vesicles was observed in the C. elegans synaptophysin/ synaptogyrin/SCAMP triple knockout (Abraham et al., 2006), in retinal rod photoreceptor cells that lacked both synaptophysin and synaptoporin (Spiwoks-­‐Becker et al., 2001), and in the squid giant synapse when the interaction between synaptophysin and dynamin was inhibited (Daly et al., 2000). Similarly, in the C. elegans synaptogyrin single knockout, synthetic phenotypes were observed with 122 endophilin, synaptojanin, and synaptotagmin, all of which participate in clathrin-­‐
mediated endocytosis (Song and Zinsmaier, 2003; Shupliakov, 2009; Abraham et al., 2011). Moreover, synaptobrevin, a known binding partner of synaptophysin (Edelmann et al., 1995), has been implicated in a rapid version of endocytosis that may not rely on clathrin-­‐coated vesicles (Deak et al., 2004). Taken together, these results suggest that synaptogyrin and/or synaptophysin may regulate a separate endocytic pathway that is independent of clathrin-­‐mediated endocytosis, e.g., kiss-­‐and-­‐run endocytosis. This raises the possibility that the lack of severe phenotypes in the synaptophysin and synaptogyrin knockouts is due to compensation by alternative endocytic mechanisms and implies that significant defects in synaptic vesicle recycling may only occur under conditions in which clathrin-­‐mediated endocytosis is insufficient to replenish synaptic vesicle pools. Analysis of the synaptophysin/synaptogyrin double knockout mouse revealed a delay in the recovery from synaptic depression induced by high-­‐frequency stimulation, again suggesting that these proteins regulate synaptic vesicle recycling kinetics (Janz et al., 1999). However, these mice have no alterations in their evoked synaptic responses or in the release of glutamate from synaptosomes. On the other hand, the double knockout displayed deficits in paired-­‐pulse facilitation (PPF), post-­‐tetanic potentiation (PTP), and long-­‐term potentiation (LTP), indicating that synaptophysin and synaptogyrin serve a functional role in synaptic plasticity (Janz et al., 1999). The observation that synaptophysin phosphorylation is enhanced in hippocampal brain slices that have undergone LTP lends additional support to the hypothesis that these proteins can modulate synaptic transmission (Mullany and Lynch, 1998). 123 We sought to further investigate the role of synaptogyrin in the regulation of synaptic vesicle exo-­‐endocytosis through the generation and characterization of a Drosophila synaptogyrin mutant. Our initial analysis revealed no obvious behavioral phenotypes, despite the fact that Drosophila has no other gyrins or physins that could compensate for the absence of synaptogyrin. The loss of synaptogyrin did not dramatically affect protein expression levels of other synaptic proteins, nor did it result in overt changes in the localization of synaptotagmin or synaptobrevin, two essential synaptic vesicle proteins. Furthermore, the absence of synaptogyrin did not impact the viability or behavior of gyr animals as determined by several different assays. We therefore proceeded to perform a more in-­‐depth analysis to search for alterations in synaptic function similar to those seen in the mouse and nematode knockouts. While gross synaptic morphology is unaffected in gyr mutants, we discovered changes in synaptic vesicle size regulation as well as alterations in the synaptic vesicle exo-­‐
endocytic cycle that are manifest under enhanced stimulation conditions. 124 Results gyr mutants have normal synaptic growth and bouton morphology As Drosophila proceed through the three larval stages, the surface area of the body wall muscles grows approximately 100-­‐fold, and synaptic innervation at the NMJ increases in parallel through the addition of new boutons to maintain proper muscle depolarization. Development of the NMJ requires coordination between the synapse and muscle and involves several trans-­‐synaptic signaling cascades including the Wnt and transforming growth factor-­‐β (TGFβ) pathways (reviewed in (Collins and DiAntonio, 2007)). Disruptions in these growth signaling pathways can lead to synaptic undergrowth, synaptic overgrowth, or defects in establishing correct synaptic architecture. Connections between the motorneuron and muscle are also highly dependent on synaptic activity, as alterations in the levels of synaptic transmission brought about by a variety of factors can dramatically influence the extent of synaptic innervation (Budnik et al., 1990). For example, increased neuronal activity induced using either temperature-­‐sensitive seizure mutants (Guan et al., 2005) or enhanced larval locomotion (Sigrist et al., 2003) can result in overgrowth of synaptic varicosities at the NMJ. Furthermore, many Drosophila endocytic and membrane recycling mutants, including rab11, endophilin, synaptojanin, and dap160, display an increase in supernumerary or “satellite” boutons (Koh et al., 2004; Dickman et al., 2006; Khodosh et al., 2006). Therefore, analysis of synaptic bouton morphology not only can provide insights into possible developmental defects, but also can hint at changes in synaptic activity or function. 125 When we examined the NMJs of third instar larvae in gyr mutants, we discovered that overall synaptic morphology is normal with no obvious synaptic overgrowth, undergrowth, or alterations in branching. The total number of synaptic boutons at muscle 6/7 is unchanged in both gyr1 and gyr2 mutant larvae compared to controls (Figure 1A-­‐B; p = 0.48, one-­‐way ANOVA). There also is no significant difference in muscle size, which is consistent with our previous observation that gyr mutants do not have developmental delays (Figure 1C; p = 0.20, one-­‐way ANOVA). Furthermore, we did not observe an increase in satellite boutons or notice any dramatic changes in bouton size or shape (data not shown). Two separate motorneurons innervate muscles 6 and 7, and while both use glutamate as their excitatory neurotransmitter, their boutons are morphologically distinct (Johansen et al., 1989; Atwood et al., 1993). Type Ib (big) boutons from the RP3 neuron are larger and are surrounded by a thicker subsynaptic reticulum (SSR) than type Is (small) boutons from the 6/7b neuron (Atwood et al., 1993; Keshishian et al., 1993). Type Ib and Is boutons can easily be distinguished by immunostaining for the protein Discs large (Dlg), a postsynaptic structural protein associated with the membrane infoldings of the SSR (Lahey et al., 1994). Since type Ib boutons are associated with a much more elaborate SSR, Dlg staining is significantly more intense at type Ib boutons. When we examined gyr third instar larvae, we found no significant difference in the number of type Ib or type Is boutons (Figure 1D-­‐E; p > 0.1, one-­‐way ANOVA), again suggesting that synaptogyrin does not play a developmental role at the larval NMJ. Similarly, the number of active zones (the sites from which synaptic vesicles are released) is also unchanged in gyr mutants compared to controls (Figure 2A-­‐B; p = 126 0.61, Student’s t-­‐test). Active zone size and spacing also appear to be unaffected in gyr mutants, and since the average number of boutons and active zones is comparable in gyr mutants and controls, the average number of active zones per bouton is almost identical (Figure 2C; control = 7.47 active zones/bouton; gyr = 7.62 active zones/ bouton). We therefore conclude that the loss of synaptogyrin in Drosophila does not significantly alter synaptic growth or influence the number of synaptic vesicle release sites. Ultrastructural analysis of gyr larvae reveals alterations in synaptic vesicle diameter and density Although bouton architecture in gyr mutants appears normal at the level of light microscopy, we performed electron microscopy at the NMJs of third instar larvae to determine whether gyr mutants have defects in ultrastructural synaptic morphology at type Ib boutons. We were surprised to discover that boutons at the larval NMJ in gyr animals display variable changes in synaptic vesicle diameter and density. In control boutons, synaptic vesicles are uniform in size and tightly clustered around the periphery of the bouton (Figure 3A-­‐B). Similarly, some gyr boutons appear indistinguishable from controls with respect to synaptic vesicle diameter and density (compare Figures 3A and 3E). However, other gyr boutons have a noticeable decrease in synaptic vesicle density (Figure 3F) and/or an increase in the number of large synaptic vesicles, the largest of which may very well be endosomes or endocytic cisternae (Figure 3C-­‐D). This dramatic increase in the variability of synaptic vesicle diameter in certain gyr boutons is illustrated in Figure 3D (compare to Figure 3B). While a fraction of gyr boutons have a marked decrease in synaptic vesicle density, 127 overall there is not a statistically significant difference in the density of synaptic vesicles per μm2 when all gyr boutons are included in the analysis (Figure 4A; p = 0.19, Student’s t-­‐test). However, there is a statistically significant increase in the mean synaptic vesicle diameter in gyr boutons compared to controls (Figure 4B; control = 42.91 ± 0.49, gyr = 45.33 ± 0.82; p = 0.02, Student’s t-­‐test). There also is an overall shift in the distribution of synaptic vesicle diameter towards larger values with an increased amount of variability in diameter, which suggests that synaptogyrin directly or indirectly regulates synaptic vesicle size (Figure 4C). As previously mentioned, muscle 6/7 is innervated by two bouton subtypes, type Ib and Is. Type Is boutons are reported to have a larger mean synaptic vesicle diameter than type Ib boutons (Karunanithi et al., 2002), which led us to question whether the increase in average synaptic vesicle diameter observed in certain gyr mutant boutons could be due to the inadvertent selection of type Is boutons instead of type Ib. We had previously determined that gyr mutants have no significant difference in the number of type Ib or Is boutons at the NMJ (Figure 1D-­‐E), nevertheless we re-­‐
examined the electron micrographs to confirm that the gyr boutons we analyzed had the thick SSRs characteristic of type Ib boutons. Furthermore, there was no significant difference in the average bouton area of the micrographs we sampled, making it unlikely that the gyr bouton analysis was biased towards smaller boutons (average bouton area in μm2 ± SEM: control = 3.21 ± 0.33, n = 24; gyr = 3.35 ± 0.36, n = 24; p = 0.76, Student’s t-­‐test). If the smaller gyr boutons were indeed type Is, we might expect that there would be a linear correlation between average synaptic vesicle diameter and the area of a 128 bouton, with smaller boutons having a larger mean vesicle diameter. However, we found only a slight, non-­‐significant trend in the gyr boutons analyzed (Figure 4D; r2 = 0.065, p = 0.23, Pearson correlation). As expected, there also was no correlation between bouton size and average synaptic vesicle diameter in control boutons (Figure 4D; r2 = 0.003, p = 0.79, Pearson correlation). Type Is boutons are also reported to have fewer mitochondria than type Ib boutons (Atwood et al., 1993), and we found no statistically significant difference between gyr mutants and controls in the number of mitochondria per μm2 (Figure 4E; p = 0.47, Student’s t-­‐test). We therefore propose that it is unlikely that the entire subgroup of gyr boutons with profoundly abnormal synaptic vesicle diameter are type Is boutons, although we cannot completely rule out the possibility that the loss of synaptogyrin alters the morphology of boutons such that a minority of type Is boutons are more similar in appearance to type Ib boutons. We conclude that the loss of synaptogyrin causes an overall increase in average synaptic vesicle diameter as well as enhanced variability of synaptic vesicle size, although the penetrance of these phenotypes is variable between individual gyr boutons. gyr mutants have an increased number of endocytic cisternae following intense stimulation Synaptic vesicles are known to form through several different pathways, including traditional clathrin-­‐mediated endocytosis from the plasma membrane (De Camilli and Takei, 1996) and via endosomal intermediates (Heuser and Reese, 1973; Takei et al., 1996; de Lange et al., 2003). During periods of relatively low activity, clathrin-­‐mediated endocytosis appears to be the predominant form of synaptic vesicle retrieval at central synapses (Granseth et al., 2006). However, intense non-­‐physiological 129 stimuli or tetanic action potential stimulation can induce bulk endocytosis, a process whereby large plasma membrane invaginations are internalized to form endocytic cisternae from which synaptic vesicles then bud (Miller and Heuser, 1984; Richards et al., 2000; Richards et al., 2003; Evans and Cousin, 2007). The presence of many large, endosomal-­‐like structures in a subset of gyr mutant boutons led us to hypothesize that synaptic vesicle recycling via endosomal intermediates might be disrupted in these boutons, resulting in a buildup of endocytic cisternae. To explore this possibility, we incubated larvae with a high-­‐potassium solution known to induce the formation of endocytic cisternae in response to the massive synaptic vesicle exocytosis caused by continuous membrane depolarization (Marxen et al., 1999; de Lange et al., 2003; Akbergenova and Bykhovskaia, 2009). We then investigated whether gyr mutants were impaired either in recovering vesicle membrane via bulk endocytosis or in resolving endocytic cisternae into synaptic vesicles. Under normal resting conditions, gyr mutants have a slight but not statistically significant increase in the number of cisternae per µm2 (here cisternae are defined as structures with a diameter greater than 80 nm; Figure 5 [left]; control = 0.84 ± 0.22; gyr = 1.36 ± 0.28; p = 0.21, Student’s t-­‐test). Immediately after a five-­‐minute incubation in high K+ (90 mM) Jan and Jan solution (Jan and Jan, 1976; Akbergenova and Bykhovskaia, 2009), both control and gyr animals show a similar increase in the number of cisternae, suggesting that gyr mutants are not impaired in this step of bulk endocytosis (Figure 5 [center]; control = 15.0 ± 0.62; gyr = 12.7 ± 0.90; p = 0.21, Student’s t-­‐test). However, when larvae are subsequently allowed to recover for ten minutes in normal (low K+) saline prior to fixation, gyr animals have a significantly 130 higher amount (~50% by area) of endocytic cisternae remaining compared to controls (Figure 5 [right]; control = 5.20 ± 0.72; gyr = 7.98 ± 0.78, p = 0.01, Student’s t-­‐test). This observation suggests that the process of resolving synaptic vesicles from endocytic cisternae is delayed in gyr mutants. To further examine synaptic vesicle recycling, we turned to the styryl dye FM1-­‐
43, which reversibly binds membranes and dramatically increases in fluorescence intensity upon membrane integration, allowing one to track compartments as they move through the exo-­‐endocytic cycle (Cochilla et al., 1999). Mutations in a variety of endocytic proteins, such as endophilin, eps15, synaptojanin, and dap160, result in decreased loading of FM1-­‐43 dye (Verstreken et al., 2002; Verstreken et al., 2003; Koh et al., 2004; Koh et al., 2007). It should be noted that mutations in these genes also can lead to a dramatic decline in synaptic vesicle number at rest (in the case of endophilin, synaptojanin, and dap160) or result in defects immediately following stimulation with high K+ (in the case of eps15), neither of which we observed in gyr mutants. When we incubated gyr and control larvae for five minutes with a high K+ (60 mM) solution containing 4 μM FM1-­‐43, we found no significant difference in the amount of dye uptake (Figure 6A-­‐B; p = 0.16, Student’s t-­‐test). This result is unsurprising given that electron micrographs clearly illustrate that gyr is capable of taking up large quantities of membrane in the form of endocytic cistenae immediately following a potassium shock (Figure 5). We then examined unloading of FM1-­‐43 dye by incubating preparations briefly (for one minute) in high K+ saline without dye. gyr and control NMJs display a similar decrease in fluorescence levels (~50%) relative to the loading levels, suggesting that an equal fraction of the dye is capable of being released in both 131 genotypes (Figure 6C; p = 0.29, Student’s t-­‐test). Since both loading and unloading were performed with a non-­‐physiological stimulus, it is difficult to draw conclusions about whether gyr mutants sequester FM1-­‐43 in endocytic cisternae to a greater extent than controls. The second potassium stimulus used to unload the dye may have caused both synaptic vesicles and cisternae to fuse, which could mask a defect in synaptic vesicle budding from cisternae. Synaptic vesicle diameter in gyr mutants shifts to more normal values following a high K+ shock If synaptogyrin is indeed involved in regulating synaptic vesicle budding from endocytic cisternae, one might expect that the dramatic synaptic vesicle recycling induced by high-­‐potassium stimulation might exacerbate the misregulation of synaptic vesicle size seen in gyr mutants under non-­‐stimulated conditions. Intriguingly, when we measured the diameter of synaptic vesicles ten minutes after high-­‐potassium stimulation, we found that the distribution of synaptic vesicle diameter was more similar to controls compared with pre-­‐potassium stimulation (Figure 7A-­‐B). The average synaptic vesicle diameter following the potassium shock was not significantly different between gyr and controls (Figure 7C; control = 42.95 ± 0.63 nm, n = 19; gyr = 43.06 ± 0.48 nm, n = 22; p = 0.89, Student’s t-­‐test), and individual gyr boutons had much less variation in their mean synaptic vesicle diameter (compare Figure 7C with Figure 4B). So while the number of large cisternae is increased in gyr mutants after intense stimulation, synaptic vesicles (less than 60 nm in diameter) that form soon after stimulation are more similar in size to controls (Figure 7D). This shift in distribution is more readily apparent when structures greater than 60 nm in diameter are excluded 132 and the cumulative frequency plots are normalized (Figure 7E). This allows for a more direct comparison of synaptic vesicle size distribution since the elevated number of endocytic cisternae after the potassium shock dramatically alters the overall cumulative distribution curves. These findings suggest that intense stimulation results in the restoration of normal synaptic vesicle diameter, perhaps by forcing the abnormally large synaptic vesicles to fuse (see Discussion). gyr mutants exhibit increased release probability and facilitation The increased synaptic vesicle diameter seen in gyr boutons lead us to hypothesize that these synapses may release more neurotransmitter per vesicle on average and might thereby result in an enhanced postsynaptic response. To our surprise, however, both miniature excitatory junctional potential (mEJP) and evoked excitatory junctional potential (EJP) amplitude were not significantly altered in gyr mutant larvae under low-­‐frequency stimulation conditions (Figure 8A-­‐B). Although the average mEJP frequency is slightly decreased in the gyr mutant compared to controls, the difference is not statistically significant (control = 3.67 ± 0.27 Hz, n = 6; gyr = 2.58 ± 0.41 Hz, n = 7; p = 0.06, Student’s t-­‐test). The lack of a significant change in mEJP amplitude suggests that either the larger synaptic vesicles and cisternae are resistant to spontaneous fusion or they release approximately the same amount of neuro-­‐
transmitter as normal-­‐sized synaptic vesicles. When we measured the diameter of synaptic vesicles located in close proximity to active zones (within 150 nm), we found no significant difference in size between gyr mutants and controls, suggesting that larger vesicles are not normally recruited to release sites (control = 41.02 ± 0.56 nm, n = 82 vesicles from 11 active zones; gyr = 41.97 ± 0.76 nm, n = 80 vesicles from 12 active 133 zones; p = 0.31, Student’s t-­‐test). This does not, however, indicate that the larger endosomal-­‐like structures are incapable of fusion under other circumstances such as during high-­‐frequency stimulation or after incubation with high-­‐potassium saline. Although basal evoked release is normal in gyr mutants, analysis using two-­‐
electrode voltage clamp at higher stimulation frequencies revealed increases in synaptic vesicle release probability. At low (0.2 mM) calcium concentrations, gyr larvae display an increase in the paired-­‐pulse ratio at short (50 ms), but not longer (100 ms) interstimulus intervals (Figure 9A). Similarly, when stimulated continuously at 10 or 20 Hz for 500 stimuli, gyr animals show an increase in facilitation that becomes more pronounced at 20 Hz (Figure 9B). Interestingly, this enhanced facilitation appears to be transient when examined using a longer stimulation protocol of 1,500 stimuli at 20 Hz. After approximately 500 stimuli, the enhanced facilitation in gyr larvae peaks and subsequently declines until facilitation reaches levels similar to that of controls (Figure 9C). Notably, this phenotype can be rescued by presynaptic expression of synaptogyrin cDNA using the pan-­‐neuronal elavc155–GAL4 promoter (Figure 9C). Interestingly, when we separated the peak current (evoked excitatory junctional current [EJC] amplitude) and the charge transferred (the total amount of current transfer induced by an action potential) during the extended 20 Hz stimulation protocol, we discovered that there was a more pronounced difference in the amount of charge transferred between gyr and control larvae than the difference in the peak EJC amplitude (Figure 9D). This suggests that, under high-­‐frequency stimulation conditions, the gyr mutant not only has more facilitation as indicated by the EJC amplitude, but also has altered release kinetics such that more current is transferred per stimulus. When 134 representative traces from gyr and control larvae are normalized, there is a clear increase in the width of the gyr EJC trace relative to the control (Figure 9E). This finding implies that gyr mutants have an increase in the amount of asynchronous synaptic vesicle release that is induced by high-­‐frequency stimulation. Unsurprisingly, the initial enhanced facilitation in gyr larvae leads to a greater total amount of charge transferred over the course of the stimulation protocol (Figure 9F); however, after the initial phase of facilitation, gyr and control larvae reach the same steady-­‐state level of release. Exactly how these changes in facilitation and release kinetics impact the overall function of the NMJ or the behavior of the animal remains to be determined. 135 Discussion Despite decades of research into the function of synaptogyrin and synaptophysin, the exact role of these proteins in the synaptic vesicle cycle remains ambiguous. We have continued our characterization of a Drosophila synaptogyrin mutant in an attempt to elucidate synaptogyrin function in a relatively simple genetic background that is free of additional synaptophysin or synaptogyrin paralogs. We examined larval NMJs using both immunofluorescence and electron microscopy and found no significant differences in neuronal architecture or synapse number. This result is unsurprising given that mouse synaptophysin/synaptogyrin double knockouts and C. elegans mutants lacking synaptophysin, synaptogyrin, and SCAMP have no significant changes in brain architecture or synaptic wiring, indicating that these proteins are not essential for nervous system development (Janz et al., 1999; Abraham et al., 2006; Abraham et al., 2011). Similarly, synaptogyrin is not required for synaptic transmission per se, as all knockouts have normal evoked responses, which demonstrates that the basic release machinery is not significantly altered by the loss of synaptogyrin (Figure 8) and (Janz et al., 1999; Abraham et al., 2006). However, further analysis of the mouse and nematode synaptophysin/synaptogyrin knockouts identified alterations in certain aspects of endocytosis, synaptic plasticity, and/or synaptic vesicle morphology. Our work with the Drosophila gyr mutant supports the idea that synaptogyrin and synaptophysin are not essential for neurotransmission, yet they play a potentially important modulatory role in the exo-­‐endocytic cycle and can impact synaptic function. 136 The relationship between synaptogyrin and synaptic activity We were intrigued to discover that a subset of gyr boutons have quite dramatic changes in vesicle morphology, while others appear indistinguishable from wild-­‐type boutons. Interestingly, work in murine retinal cells lacking both synaptophysin and its paralog synaptoporin indicated that synaptic vesicle density was decreased and that this phenotype became more pronounced during periods of high activity (i.e., during dark adaptation when photoreceptors release more synaptic vesicles). Moreover, during these periods of elevated activity there was an increase in the number of large (>80 nm) vacuolar-­‐like structures as well as an increase in synaptic vesicle diameter in synaptophysin knockouts compared to controls (Spiwoks-­‐Becker et al., 2001). These results suggest that synaptophysin promotes the efficient formation of synaptic vesicles and that enhanced activity can reveal more severe phenotypes. Our work indicates that synaptogyrin performs a similar role in modulating synaptic vesicle formation in Drosophila, and it is possible that the phenotypic variation we see at gyr boutons under resting conditions (Figure 3) could be due to dissimilar levels of activity between boutons. Several studies have shown that bouton activity levels are variable at the Drosophila NMJ, namely that terminal boutons tend to have higher release probabilities than more proximal boutons (Guerrero et al., 2005; Peled and Isacoff, 2011). Interestingly, there also was a great deal of variability in the severity of the phenotypes in murine retinal photoreceptor cells lacking both synaptophysin and synaptoporin (Spiwoks-­‐Becker et al., 2001), suggesting that synaptophysin and/or synaptogyrin may not be required to the same extent in every nerve terminal. We were unable to retrospectively determine the relative positions of the boutons we examined 137 by electron microscopy to correlate the severity of phenotype with the location of a bouton along the synaptic arbor. However, it is possible to mark boutons prior to preparation for electron microscopy (e.g., with a cactus needle), so future studies could examine the relationship between synaptic activity and synaptic vesicle morphology. While it may be the case that a more active bouton would require synaptogyrin for proper vesicle formation (and would therefore have a more severe phenotype), the finding that a high-­‐potassium shock causes gyr synaptic vesicle diameter to more closely resemble controls suggests that the opposite may be true. In other words, it appears as though higher activity leads to more normal synaptic vesicle formation relative to controls, at least for those vesicles less than 60 nm in diameter. However, since gyr boutons have an increased number of cisternae following high-­‐potassium stimulation (Figure 5), it is clear that dramatic increases in neuronal activity can cause alterations in endocytic pathways in gyr mutants, which subsequently leads to an accumulation of abnormally large structures. A preliminary model to explain synaptogyrin’s role in synaptic vesicle formation While there are still many unanswered questions about synaptogyrin function, we have developed a preliminary model to account for the gyr phenotypes related to abnormal synaptic vesicle diameter and the delay in the recovery of endocytic cisternae. Numerous studies have established the existence of multiple endocytic pathways at the NMJ, including clathrin-­‐mediated and bulk endocytosis, and, according to some, the somewhat controversial kiss-­‐and-­‐run endocytosis (Koenig and Ikeda, 1996; Verstreken et al., 2002; Dickman et al., 2005). We propose that there are at least two pathways responsible for the budding of synaptic vesicles from endocytic cisternae, 138 one of which is regulated by synaptogyrin and the other that is independent of synaptogyrin function. In a wild-­‐type bouton, both pathways are fully functional and are able to produce vesicles of normal size, resulting in a low level of abnormal synaptic vesicles or structures resembling endocytic cisternae. However, in the absence of synaptogyrin, the synaptogyrin-­‐dependent pathway can no longer function properly and produces vesicles of abnormal size. The synaptogyrin-­‐independent pathway is still functional and capable of generating synaptic vesicles with a normal diameter; nevertheless, there is an overall increase in synaptic vesicle diameter in gyr animals. Since these larger synaptic vesicles appear to be excluded from active zones (see Results), it is possible that they are incapable of exocytosis under basal conditions and therefore accumulate over time. During periods of intense activity that induce bulk endocytosis, a wild-­‐type synapse would have both pathways fully capable of regenerating synaptic vesicles from endocytic cisternae, and these structures quickly would resolve into synaptic vesicles. However, in the gyr mutant, the synaptogyrin-­‐dependent pathway is impaired, and the time it takes to resolve cisternae into synaptic vesicles would be increased relative to controls. Furthermore, if the synaptogyrin-­‐independent pathway is dominant in the gyr mutant, the synaptic vesicles that form shortly after the high-­‐potassium shock would be more uniform in size because the vast majority would have been generated through the unimpaired synaptogyrin-­‐independent pathway. We anticipate that, if this model were true, as the gyr boutons recover from a high-­‐potassium shock, the impairment of the synaptogyrin-­‐dependent pathway would result in the accumulation of abnormally sized vesicles, similar to what we observed under resting conditions. 139 An alternative possibility to the multiple endocytic pathways described above would be that a single major endocytic pathway is responsible for vesicle biogenesis from cisternae, and the loss of synaptogyrin merely slows the process of synaptic vesicle budding. While this certainly would explain the delay in resolving endocytic cisternae into synaptic vesicles, it is unclear how this model would account for the changes in synaptic vesicle diameter seen following a potassium shock. If synaptogyrin were involved in the budding of all synaptic vesicles from endocytic cisternae, one would expect to see a population of abnormally shaped synaptic vesicles shortly after the high-­‐potassium shock as the cisternae begin to resolve, which we did not observe. However, it is unknown what fraction of vesicles generated after the potassium shock originate directly from the plasma membrane as opposed to cisternae. If the majority of synaptic vesicles observed ten minutes after the potassium shock derive from the plasma membrane and this form of endocytosis is independent of synaptogyrin, then the plasma membrane-­‐derived synaptic vesicles may mask a smaller population of abnormal synaptic vesicles originating from cisternae in the gyr mutant. This also raises the need to determine the extent to which synaptogyrin mediates endocytosis from the plasma membrane. In future studies, it will be informative to track the flow of synaptic vesicle and cisternal membranes during various stimulation conditions using a combination of photoconvertible FM dyes and electron microscopy, which may allow us to more precisely determine synaptogyrin’s influence on synaptic vesicle recycling. Synaptogyrin influences synaptic vesicle release probability The lack of significant changes in miniature or evoked EJP amplitude suggests that the abnormally large synaptic vesicles and endosomal-­‐like cisternae seen in gyr 140 boutons are not capable of fusing under basal conditions. However, the increase in the paired-­‐pulse ratio suggests that synaptic vesicle release probability is enhanced in gyr animals. Furthermore, vesicle fusion dynamics are altered during high-­‐frequency stimulation, resulting in enhanced facilitation and a greater amount of total synaptic vesicle release. Some of the enhanced facilitation seen in gyr mutants may be due to the recruitment of larger synaptic vesicles during the initial phases of high-­‐frequency stimulation, and the lack of sustained facilitation would be due to the depletion of these larger synaptic vesicles. However, the finding that increased facilitation occurs in the gyr mutant by the second action potential (as illustrated by the paired-­‐pulse ratio) suggests that this explanation is insufficient, as the larger synaptic vesicles would have to be trafficking to active zones exceptionally quickly to account for this rapid facilitation. How might synaptogyrin regulate synaptic vesicle release? One possibility is that synaptogyrin mediates exocytosis through a direct interaction with a component of the vesicle fusion machinery such as synaptobrevin. Although synaptobrevin and synaptogyrin have not been shown to interact directly, synaptobrevin does bind synaptophysin, and synaptophysin has been proposed to regulate synaptobrevin availability in a manner similar to the syntaxin/Munc18 interaction (Hata et al., 1993; Calakos and Scheller, 1994; Edelmann et al., 1995; Washbourne et al., 1995). Synaptophysin and/or synaptogyrin may control the ability of synaptobrevin to interact with the other members of the neuronal SNARE complex and may thereby ensure that the SNARE complex only assembles under appropriate circumstances. The loss of 141 synaptophysin or synaptogyrin could therefore result in inappropriate SNARE complex formation, which could alter synaptic vesicle release properties. Alternatively, the changes in synaptic vesicle fusion seen in gyr mutants may instead be a consequence of the defects in synaptic vesicle formation that we observed by electron microscopy. If vesicle biogenesis is impaired in the absence of synaptogyrin, one of the consequences may be that synaptic vesicles do not receive the appropriate complement of synaptic vesicle proteins, which could lead to alterations in the dynamics of synaptic vesicle exocytosis. A connection between changes in synaptic function and protein sorting via endosomal compartments was recently described in a paper investigating the function of a neuronal Rab35 GTPase activating protein (GAP) named Skywalker (Sky) that is proposed to regulate synaptic vesicle trafficking between endosomal compartments (Uytterhoeven et al., 2011). Sky appears to inhibit the trafficking of synaptic vesicles to sorting endosomes and consequently impacts the degradation of synaptic vesicle proteins. Like gyr mutants, sky larvae accumulate an excess of cisternal-­‐like structures after intense stimulation and show increased neurotransmitter release relative to controls. Unlike gyr mutants, however, sky larvae have a basal increase in exocytosis and also display punctate accumulations of FM1-­‐43 after high-­‐potassium stimulation that we did not observe in gyr larvae (data not shown). Nevertheless, the sky phenotypes demonstrate that changes in synaptic vesicle protein sorting through endosomal compartments can lead to alterations in synaptic vesicle release properties. It remains to be determined whether synaptogyrin might also directly or indirectly regulate the protein composition of synaptic vesicles. 142 Methods Fly stocks and transgenics Flies were cultured using standard media and techniques at room temperature (~22oC) unless otherwise noted. The GAL4/UAS system was utilized to drive neuronal expression of selected transgenes (Brand and Perrimon, 1993). The UAS-­‐myc-­‐gyrin construct used for rescue experiments was subcloned into a modified pValum vector with an N-­‐terminal myc tag (Cho et al., 2010). This construct was injected into the strain yv;;attP2, which contains a site for targeted transgene insertion on the third chromosome (Markstein et al., 2008; Ni et al., 2008). The pan-­‐neuronal driver elavc155-­
GAL4 was used to express the transgene in neurons. Embryo injections were performed at Genetic Services, Inc. (Cambridge, MA). Immunohistochemistry Wandering third instar larvae were dissected in calcium-­‐free HL3.1 saline (70 mM NaCl, 5 mM KCl, 10 mM NaHCO3, 4 mM MgCl2, 5 mM trehalose, 115 mM sucrose, 5 mM HEPES, pH 7.2) and fixed for 45 minutes in HL3.1 containing 4% formaldehyde. Following several washes in phosphate-­‐buffered saline (PBS) and PBST (PBS with 0.1% Tween 20), larvae were incubated with primary antibody in PBST overnight at 4oC. After additional washes, larvae were incubated in secondary antibodies in PBST for four hours at room temperature, washed, and mounted in 70% glycerol. The dilutions for primary antibodies were: nc82/bruchpilot, 1:100; Dlg, 1:250. The nc82 antibody developed by Erich Buchner and the Dlg (4F3) antibody developed by Corey Goodman were obtained from the Developmental Studies Hybridoma Bank developed under the 143 auspices of the NICHD and maintained by the University of Iowa, Department of Biological Sciences, Iowa City, IA 52242. Secondary antibodies were used at a dilution of 1:250 and include Alexa Fluor 488-­‐conjugated goat anti-­‐mouse (Invitrogen) and Rhodamine Red-­‐conjugated donkey anti-­‐mouse (Jackson ImmunoResearch). Goat α-­‐
HRP antibodies conjugated to DyLight 549 (Jackson ImmunoResearch) were added with secondary antibodies and were used at a concentration of 1:500. Images were acquired using a confocal microscope (Axoplan 2; Carl Zeiss MicroImaging, Inc.) running PASCAL software (Carl Zeiss) with a 40x oil-­‐immersion lens. Bouton and active zone counting Age-­‐matched larvae were grown at low density at 25oC and immunohisto-­‐
chemistry was performed as described above. Bouton quantification was performed using antibodies against horseradish peroxidase (HRP), a neuronal membrane marker, and Discs large (Dlg), a postsynaptic scaffolding protein. Type Ib and Is boutons were differentiated by the intensity of Dlg immunofluorescence, since type Ib boutons have significantly higher levels of Dlg (Lahey et al., 1994). Active zone numbers were quantified using monoclonal nc82 antibodies, which recognize the active zone component bruchpilot (Wagh et al., 2006). To control for variability in staining as much as possible, larvae of different genotypes were combined and processed together (tails were cut to differentiate genotypes). All bouton and active zone measurements were done at muscle 6/7 of segment A3. Imaging was performed using an Axoplan 2 confocal microscope with PASCAL software (Carl Zeiss MicroImaging, Inc.). Image analysis was done using ImageJ software (NIH). Bouton number is known to be proportional to 144 muscle size (Lnenicka and Keshishian, 2000); however, since muscle area did not differ significantly between gyr and control larvae, no corrections were necessary (see Results). Bouton and active zone counting were performed blind to genotype. Electron microscopy Dissected Drosophila third instar larvae were fixed for 1 hour in 4% para-­‐
formaldehyde/1% glutaraldehyde in 0.1 M cacodylate buffer (pH 7.2). Fixed larvae were then washed in standard HL3 solution for 1 hour and postfixed for 40 minutes in 1% osmium tetroxide. Specimens were then dehydrated in a graded series of ethanol and water mixtures up to 100% ethanol followed by acetone. Samples were pre-­‐
infiltrated for 1 hour in a 1:1 mixture of acetone and Epon and then embedded in fresh Epon overnight at 60oC. Thin sections (60-­‐70 nm) were contrasted with 0.2% lead citrate and imaged at 80 kV on an FEW Tecnai G2 Spirit transmission electron microscope equipped with an AMT CCD camera. Imaging was performed at the W.M. Keck Microscopy Facility at the Whitehead Institute, Cambridge, MA. Quantification of synaptic vesicle diameter and density was performed using randomly sampled images from type Ib boutons. The diameter of synaptic vesicles and cisternae were measured from the outsides of the vesicle membranes along the long axis. A vesicle was measured only if its borders were well defined and it had a clear center (i.e., it was not a dense-­‐core vesicle). Structures larger than 60 nm in diameter were excluded from calculations of mean synaptic vesicle diameter, but all vesicles that could be measured were included in the frequency and cumulative frequency analysis. High-­‐potassium stimulation was performed using a modified Jan and Jan solution (Jan and Jan, 1976; 45 mM NaCl, 90 mM KCl, 36 mM sucrose, 2 mM CaCl2, 2 mM MgCl2, 5 mM 145 HEPES, pH 7.3). Cisternae were defined as intracellular single-­‐membrane structures with a diameter greater than or equal to 80 nm. In certain cases the exact size of a cisterna could not be measured; therefore, it was not included in the cumulative frequency plots (Figure 6). However, if a cisterna clearly was greater than 80 nm in diameter, it was included in the analysis of the number of cisternae per unit area (Figure 5). The analysis was done blind to genotype wherever possible. Measurements were performed using ImageJ (NIH) and Adobe Photoshop 7.0 (Adobe Systems, Inc.) software. Electrophysiology Intracellular electrophysiology was performed on wandering third instar larvae at room temperature as described (Rieckhof et al., 2003) at muscle fiber 6 of segments A3 and A4 using HL3.1 saline with 0.2 mM CaCl2. Data acquisition and analysis were performed using an Axoclamp 2B amplifier and pClamp 9.0 software (Axon Instruments, Foster City, CA). mEJP amplitude was measured using Mini Analysis Program (Synaptosoft). The resting muscle potential was not significantly different between genotypes in the traces used for mEJP analysis (control = -­‐67.4 ± 2.30 mV, n = 6; gyr = -­‐66.9 ± 1.90 mV, n = 7; p = 0.85, Student’s t-­‐test). Voltage clamp electrophysiology was performed as described (Acharya et al., 1998) using two-­‐microelectrode voltage clamp (OC725, Warner Instruments, Hamden, CT) at -­‐80 mV holding potential using a programmable stimulator (Master-­‐8, A.M.P.I., Jerusalem, Israel). Data acquisition and analysis were performed using pClamp 9.0 software (Axon Instruments, Foster City, CA). 146 FM1-­43 Wandering third instar larvae were dissected in HL3.1 solution with 1 mM CaCl2 and the nerves were cut to prevent spontaneous muscle contractions. Larvae were stimulated for five minutes with high-­‐potassium (60 mM) HL3.1 containing 4 μM FM1-­‐
43 (Invitrogen). Preparations were briefly rinsed in standard HL3.1 and momentarily incubated in HL3.1 containing 100 μM Advasep-­‐7 (Biotium), which has been shown to reduce background staining (Kay et al., 1999). Larvae were briefly washed again in HL3.1 and then imaged to quantify the level of FM1-­‐43 uptake. After ten minutes, boutons were unloaded with a one-­‐minute incubation with high-­‐potassium (60 mM) HL3.1, briefly washed in normal HL3.1, and then imaged again. Images were acquired using an UltraVIEW VoX confocal imaging system (PerkinElmer) equipped with an ImagEM camera (C9100-­‐13, Hamamatsu) and a Yokogawa CSU-­‐X1 spinning disk head. Images were acquired with a 63x water-­‐immersion lens, and analysis was performed using Volocity 3D Image Analysis Software (PerkinElmer). Graphing and statistics Statistical analysis was performed using GraphPad Prism for Mac OS X version 5.0a. Error measurements are standard error of the mean (SEM). Statistical significance was determined using Student’s t-­‐tests, one-­‐way ANOVA, or Pearson’s correlation coefficient (*, p < 0.05; **, p < 0.01). 147 References Abraham C, Bai L, Leube RE (2011) Synaptogyrin-­‐dependent modulation of synaptic neurotransmission in Caenorhabditis elegans. Neuroscience 190:75-­‐88. Abraham C, Hutter H, Palfreyman MT, Spatkowski G, Weimer RM, Windoffer R, Jorgensen EM, Leube RE (2006) Synaptic tetraspan vesicle membrane proteins are conserved but not needed for synaptogenesis and neuronal function in Caenorhabditis elegans. Proc Natl Acad Sci U S A 103:8227-­‐8232. Acharya JK, Labarca P, Delgado R, Jalink K, Zuker CS (1998) Synaptic defects and compensatory regulation of inositol metabolism in inositol polyphosphate 1-­‐
phosphatase mutants. Neuron 20:1219-­‐1229. Akbergenova Y, Bykhovskaia M (2009) Enhancement of the endosomal endocytic pathway increases quantal size. Mol Cell Neurosci 40:199-­‐206. Atwood HL, Govind CK, Wu CF (1993) Differential ultrastructure of synaptic terminals on ventral longitudinal abdominal muscles in Drosophila larvae. J Neurobiol 24:1008-­‐1024. Belizaire R, Komanduri C, Wooten K, Chen M, Thaller C, Janz R (2004) Characterization of synaptogyrin 3 as a new synaptic vesicle protein. J Comp Neurol 470:266-­‐281. Brand AH, Perrimon N (1993) Targeted gene expression as a means of altering cell fates and generating dominant phenotypes. Development 118:401-­‐415. Budnik V, Zhong Y, Wu CF (1990) Morphological plasticity of motor axons in Drosophila mutants with altered excitability. J Neurosci 10:3754-­‐3768. Calakos N, Scheller RH (1994) Vesicle-­‐associated membrane protein and synaptophysin are associated on the synaptic vesicle. J Biol Chem 269:24534-­‐24537. Cho RW, Song Y, Littleton JT (2010) Comparative analysis of Drosophila and mammalian complexins as fusion clamps and facilitators of neurotransmitter release. Mol Cell Neurosci 45:389-­‐397. Cochilla AJ, Angleson JK, Betz WJ (1999) Monitoring secretory membrane with FM1-­‐43 fluorescence. Annu Rev Neurosci 22:1-­‐10. Collins CA, DiAntonio A (2007) Synaptic development: insights from Drosophila. Curr Opin Neurobiol 17:35-­‐42. Daly C, Sugimori M, Moreira JE, Ziff EB, Llinas R (2000) Synaptophysin regulates clathrin-­‐independent endocytosis of synaptic vesicles. Proc Natl Acad Sci U S A 97:6120-­‐6125. De Camilli P, Takei K (1996) Molecular mechanisms in synaptic vesicle endocytosis and recycling. Neuron 16:481-­‐486. de Lange RP, de Roos AD, Borst JG (2003) Two modes of vesicle recycling in the rat calyx of Held. J Neurosci 23:10164-­‐10173. Deak F, Schoch S, Liu X, Sudhof TC, Kavalali ET (2004) Synaptobrevin is essential for fast synaptic-­‐vesicle endocytosis. Nat Cell Biol 6:1102-­‐1108. Dickman DK, Horne JA, Meinertzhagen IA, Schwarz TL (2005) A slowed classical pathway rather than kiss-­‐and-­‐run mediates endocytosis at synapses lacking synaptojanin and endophilin. Cell 123:521-­‐533. Dickman DK, Lu Z, Meinertzhagen IA, Schwarz TL (2006) Altered synaptic development and active zone spacing in endocytosis mutants. Curr Biol 16:591-­‐598. 148 Edelmann L, Hanson PI, Chapman ER, Jahn R (1995) Synaptobrevin binding to synaptophysin: a potential mechanism for controlling the exocytotic fusion machine. EMBO J 14:224-­‐231. Evans GJ, Cousin MA (2007) Activity-­‐dependent control of slow synaptic vesicle endocytosis by cyclin-­‐dependent kinase 5. J Neurosci 27:401-­‐411. Granseth B, Odermatt B, Royle SJ, Lagnado L (2006) Clathrin-­‐mediated endocytosis is the dominant mechanism of vesicle retrieval at hippocampal synapses. Neuron 51:773-­‐786. Guan Z, Saraswati S, Adolfsen B, Littleton JT (2005) Genome-­‐wide transcriptional changes associated with enhanced activity in the Drosophila nervous system. Neuron 48:91-­‐107. Guerrero G, Reiff DF, Agarwal G, Ball RW, Borst A, Goodman CS, Isacoff EY (2005) Heterogeneity in synaptic transmission along a Drosophila larval motor axon. Nat Neurosci 8:1188-­‐1196. Hata Y, Slaughter CA, Sudhof TC (1993) Synaptic vesicle fusion complex contains unc-­‐18 homologue bound to syntaxin. Nature 366:347-­‐351. Heuser JE, Reese TS (1973) Evidence for recycling of synaptic vesicle membrane during transmitter release at the frog neuromuscular junction. J Cell Biol 57:315-­‐344. Hubner K, Windoffer R, Hutter H, Leube RE (2002) Tetraspan vesicle membrane proteins: synthesis, subcellular localization, and functional properties. Int Rev Cytol 214:103-­‐159. Jan LY, Jan YN (1976) L-­‐glutamate as an excitatory transmitter at the Drosophila larval neuromuscular junction. J Physiol 262:215-­‐236. Janz R, Sudhof TC, Hammer RE, Unni V, Siegelbaum SA, Bolshakov VY (1999) Essential roles in synaptic plasticity for synaptogyrin I and synaptophysin I. Neuron 24:687-­‐700. Johansen J, Halpern ME, Johansen KM, Keshishian H (1989) Stereotypic morphology of glutamatergic synapses on identified muscle cells of Drosophila larvae. J Neurosci 9:710-­‐725. Karunanithi S, Marin L, Wong K, Atwood HL (2002) Quantal size and variation determined by vesicle size in normal and mutant Drosophila glutamatergic synapses. J Neurosci 22:10267-­‐10276. Kay AR, Alfonso A, Alford S, Cline HT, Holgado AM, Sakmann B, Snitsarev VA, Stricker TP, Takahashi M, Wu LG (1999) Imaging synaptic activity in intact brain and slices with FM1-­‐43 in C. elegans, lamprey, and rat. Neuron 24:809-­‐817. Keshishian H, Chiba A, Chang TN, Halfon MS, Harkins EW, Jarecki J, Wang L, Anderson M, Cash S, Halpern ME, et al. (1993) Cellular mechanisms governing synaptic development in Drosophila melanogaster. J Neurobiol 24:757-­‐787. Khodosh R, Augsburger A, Schwarz TL, Garrity PA (2006) Bchs, a BEACH domain protein, antagonizes Rab11 in synapse morphogenesis and other developmental events. Development 133:4655-­‐4665. Koenig JH, Ikeda K (1996) Synaptic vesicles have two distinct recycling pathways. J Cell Biol 135:797-­‐808. Koh TW, Verstreken P, Bellen HJ (2004) Dap160/intersectin acts as a stabilizing scaffold required for synaptic development and vesicle endocytosis. Neuron 43:193-­‐205. 149 Koh TW, Korolchuk VI, Wairkar YP, Jiao W, Evergren E, Pan H, Zhou Y, Venken KJ, Shupliakov O, Robinson IM, O'Kane CJ, Bellen HJ (2007) Eps15 and Dap160 control synaptic vesicle membrane retrieval and synapse development. J Cell Biol 178:309-­‐322. Lahey T, Gorczyca M, Jia XX, Budnik V (1994) The Drosophila tumor suppressor gene dlg is required for normal synaptic bouton structure. Neuron 13:823-­‐835. Lnenicka GA, Keshishian H (2000) Identified motor terminals in Drosophila larvae show distinct differences in morphology and physiology. J Neurobiol 43:186-­‐197. Markstein M, Pitsouli C, Villalta C, Celniker SE, Perrimon N (2008) Exploiting position effects and the gypsy retrovirus insulator to engineer precisely expressed transgenes. Nat Genet 40:476-­‐483. Marxen M, Volknandt W, Zimmermann H (1999) Endocytic vacuoles formed following a short pulse of K+ -­‐stimulation contain a plethora of presynaptic membrane proteins. Neuroscience 94:985-­‐996. Miller TM, Heuser JE (1984) Endocytosis of synaptic vesicle membrane at the frog neuromuscular junction. J Cell Biol 98:685-­‐698. Mullany PM, Lynch MA (1998) Evidence for a role for synaptophysin in expression of long-­‐term potentiation in rat dentate gyrus. Neuroreport 9:2489-­‐2494. Mutch SA, Kensel-­‐Hammes P, Gadd JC, Fujimoto BS, Allen RW, Schiro PG, Lorenz RM, Kuyper CL, Kuo JS, Bajjalieh SM, Chiu DT (2011) Protein quantification at the single vesicle level reveals that a subset of synaptic vesicle proteins are trafficked with high precision. J Neurosci 31:1461-­‐1470. Ni JQ, Markstein M, Binari R, Pfeiffer B, Liu LP, Villalta C, Booker M, Perkins L, Perrimon N (2008) Vector and parameters for targeted transgenic RNA interference in Drosophila melanogaster. Nat Methods 5:49-­‐51. Peled ES, Isacoff EY (2011) Optical quantal analysis of synaptic transmission in wild-­‐
type and rab3-­‐mutant Drosophila motor axons. Nat Neurosci 14:519-­‐526. Richards DA, Guatimosim C, Betz WJ (2000) Two endocytic recycling routes selectively fill two vesicle pools in frog motor nerve terminals. Neuron 27:551-­‐559. Richards DA, Guatimosim C, Rizzoli SO, Betz WJ (2003) Synaptic vesicle pools at the frog neuromuscular junction. Neuron 39:529-­‐541. Rieckhof GE, Yoshihara M, Guan Z, Littleton JT (2003) Presynaptic N-­‐type calcium channels regulate synaptic growth. J Biol Chem 278:41099-­‐41108. Shupliakov O (2009) The synaptic vesicle cluster: a source of endocytic proteins during neurotransmitter release. Neuroscience 158:204-­‐210. Sigrist SJ, Reiff DF, Thiel PR, Steinert JR, Schuster CM (2003) Experience-­‐dependent strengthening of Drosophila neuromuscular junctions. J Neurosci 23:6546-­‐6556. Song W, Zinsmaier KE (2003) Endophilin and synaptojanin hook up to promote synaptic vesicle endocytosis. Neuron 40:665-­‐667. Spiwoks-­‐Becker I, Vollrath L, Seeliger MW, Jaissle G, Eshkind LG, Leube RE (2001) Synaptic vesicle alterations in rod photoreceptors of synaptophysin-­‐deficient mice. Neuroscience 107:127-­‐142. Takamori S et al. (2006) Molecular anatomy of a trafficking organelle. Cell 127:831-­‐846. Takei K, Mundigl O, Daniell L, De Camilli P (1996) The synaptic vesicle cycle: a single vesicle budding step involving clathrin and dynamin. J Cell Biol 133:1237-­‐1250. 150 Uytterhoeven V, Kuenen S, Kasprowicz J, Miskiewicz K, Verstreken P (2011) Loss of skywalker reveals synaptic endosomes as sorting stations for synaptic vesicle proteins. Cell 145:117-­‐132. Verstreken P, Kjaerulff O, Lloyd TE, Atkinson R, Zhou Y, Meinertzhagen IA, Bellen HJ (2002) Endophilin mutations block clathrin-­‐mediated endocytosis but not neurotransmitter release. Cell 109:101-­‐112. Verstreken P, Koh TW, Schulze KL, Zhai RG, Hiesinger PR, Zhou Y, Mehta SQ, Cao Y, Roos J, Bellen HJ (2003) Synaptojanin is recruited by endophilin to promote synaptic vesicle uncoating. Neuron 40:733-­‐748. Wagh DA, Rasse TM, Asan E, Hofbauer A, Schwenkert I, Durrbeck H, Buchner S, Dabauvalle MC, Schmidt M, Qin G, Wichmann C, Kittel R, Sigrist SJ, Buchner E (2006) Bruchpilot, a protein with homology to ELKS/CAST, is required for structural integrity and function of synaptic active zones in Drosophila. Neuron 49:833-­‐844. Washbourne P, Schiavo G, Montecucco C (1995) Vesicle-­‐associated membrane protein-­‐
2 (synaptobrevin-­‐2) forms a complex with synaptophysin. Biochem J 305 ( Pt 3):721-­‐724. 151 152 Figure 1. The number of synaptic varicosities is unchanged in gyr mutants. A) Synaptic bouton number and morphology at muscle 6/7 in body segment A3 appear normal in gyr1 and gyr2 third instar larvae relative to controls (gyrPE), as determined by immunostaining against HRP (magenta) and Dlg (green). Dlg thickness is increased at type Ib boutons. B) Quantification of the total bouton number at muscle 6/7 revealed no significant difference between control, gyr1, and gyr2 (p = 0.48, one-­‐way ANOVA). Average total bouton number ± SEM: control = 70.8 ± 2.62, n = 23; gyr1 = 66.9 ± 2.80, n = 22; gyr2 = 65.2 ± 4.45, n = 21. C) The average muscle area at segment A3 muscle 6/7 is not significantly different between controls and gyr mutants (p = 0.20, one-­‐way ANOVA). Average muscle area (in μm2) ± SEM: control = 56,301 ± 1,832, n = 22; gyr1 = 61,337 ± 2,172, n = 23; gyr2 = 59,433 ± 1,976, n = 20. D) The number of type Ib boutons (as defined by thick Dlg immunofluorescence) is not significantly different between genotypes tested (p = 0.50, one-­‐way ANOVA). Average type Ib bouton number ± SEM: control = 28.3 ± 1.10, n = 23; gyr1 = 27.4 ± 0.80, n = 22; gyr2 = 25.9 ± 2.13, n = 21. E) Type Is bouton number also is unchanged in gyr mutants relative to controls (p = 0.41, one-­‐way ANOVA). Average type Is bouton number ± SEM: control = 42.6 ± 2.09, n = 23; gyr1 = 39.5 ± 2.33, n = 22; gyr2 = 39.3 ± 4.01, n = 21. Error bars are ± SEM. Scale bar = 40 μm. 153 154 Figure 2. Quantification of active zone number at gyr NMJs. A) Representative images of control and gyr NMJs stained with the monoclonal antibody nc82 to label active zones. Boxed areas are shown at higher magnification in the panels on the right. B) Active zone number is not significantly different in gyr mutants compared to controls (p = 0.61, Student’s t-­‐test). Average active zone number ± SEM: control = 529 ± 24.4, n = 14; gyr = 510 ± 28.7, n = 14. C) The number of active zones per bouton is similar in gyr mutants (7.62 active zones/bouton) relative to controls (7.47 active zones/bouton). Scale bars in (A): left = 20 μm; right = 2.5 μm. 155 156 Figure 3. Ultrastructural analysis reveals alterations in synaptic vesicle diameter in gyr mutants. A) A representative image of a synaptic bouton from a control larva. B) The boxed region in (A) shown at higher magnification to illustrate the homogeneity of synaptic vesicle diameter in control animals. C) Example of a gyr bouton with abnormal synaptic vesicle diameter and large cisternal-­‐like structures (indicated by an asterisk). D) The boxed region in (C) is shown at higher magnification to exemplify the variability in synaptic vesicle diameter (see asterisk). E) A representative gyr bouton with similar synaptic vesicle diameter and density relative to controls. F) Example of a gyr bouton with decreased synaptic vesicle density. Overall, there is not a statistically significant difference in synaptic vesicle density between gyr mutants and controls (see Figure 4; p = 0.19, Student’s t-­‐test). Scale bars: A, C, E, and F = 1 μm; B and D = 100 nm. 157 158 Figure 4. Quantification of ultrastructural morphology. A) Synaptic vesicle density is not significantly different in gyr mutants compared to controls (p = 0.19, Student’s t-­‐
test). Each point on the graph represents the synaptic vesicle density of a single bouton. Average synaptic vesicle density per bouton area (μm2) ± SEM: control = 132.6 ± 13.8, n = 16; gyr = 111.4 ± 9.3, n = 24. B) The average synaptic vesicle diameter (vesicles < 60 nm in diameter) is increased in gyr mutants relative to controls (p = 0.02, Student’s t-­‐
test). Each data point represents the mean synaptic vesicle diameter of a single bouton. At least 30 synaptic vesicles were measured in each bouton to obtain the mean synaptic vesicle diameter. In total, 2,518 synaptic vesicles from control boutons were measured, and 1,535 synaptic vesicles from gyr mutants were measured. Average synaptic vesicle diameter (in nm) ± SEM: control = 42.91 ± 0.49, n = 24; gyr = 45.33 ± 0.82, n = 24. C) Synaptic vesicle diameter is shifted to higher values relative to the control in gyr mutant larvae. Since varying numbers of synaptic vesicles were measured in each bouton, the frequency distribution of each bouton was calculated and normalized to obtain the overall frequency distribution. The measurements are from the same data set as in (B). D) Examination of the relationship between bouton area and mean synaptic vesicle diameter. Each bouton used to determine the average synaptic vesicle diameter is plotted with the bouton area along the abscissa and the mean synaptic vesicle diameter along the ordinate. The horizontal dotted lines represent the overall average for control (black) and gyr (red). The solid lines indicate the linear regression best-­‐fit values. The slopes are not significantly different from zero for either control (p = 0.79, Pearson’s correlation) or gyr (p = 0.23). E) The average number of mitochondria per bouton area is not significantly different between gyr and control larvae (p = 0.47, Student’s t-­‐test). Average number of mitochondria per μm2 ± SEM: control = 0.77 ± 0.13, n = 24; gyr = 0.95 ± 0.29, n = 28. 159 160 Figure 5. gyr mutants are delayed in resolving endocytic cisternae following a high-­‐
potassium shock. Left panels: At rest, gyr mutants have a slight, but not statistically significant, increase in the number of cisternae (endosomal-­‐like structures larger than 80 nm in diameter) per μm2 compared to controls (p = 0.21, Student’s t-­‐test). Middle panels: Immediately following a five-­‐minute incubation with high K+ (90 mM) Jan and Jan solution, control and gyr boutons have a similar dramatic increase in the number of endocytic cisternae due to bulk endocytosis (p = 0.21, Student’s t-­‐test). Right panels: Ten minutes after high K+ stimulation, gyr boutons have ~50% more endocytic cisternae than controls, suggesting that they cannot resolve endocytic cisternae into synaptic vesicles as quickly as controls (p = 0.01, Student’s t-­‐test). Average number of cisternae per μm2 ± SEM: At rest, control = 0.84 ± 0.22, n = 16; gyr = 1.36 ± 0.28, n = 29; High K+, control = 15.0 ± 0.62, n = 8; gyr = 12.7 ± 0.90, n = 5; 10’ recovery, control = 5.20 ± 0.72, n = 24; gyr = 7.98 ± 0.78, n = 23. Scale bar = 500 nm. 161 162 Figure 6. gyr mutants are not defective in loading or unloading FM1-­‐43. A) Representative images of control (top) and gyr (bottom) NMJs from FM1-­‐43 uptake experiments. Both control and gyr boutons display prominent fluorescence in synaptic boutons following incubation with a high K+ (60 mM) HL3 solution containing FM1-­‐43 (4 μM, left panels). After extensive washing to remove excess dye, terminals were unloaded with a one-­‐minute stimulation with high K+ solution in the absence of dye (right panels). B) Quantification of the FM1-­‐43 uptake experiment reveals no significant difference in mean fluorescence intensity between gyr and control boutons (p = 0.16, Student’s t-­‐test). Mean fluorescence intensity (arbitrary units) ± SEM: control = 6,143 ± 473, n = 9; gyr = 5,222 ± 417, n = 9. C) Control and gyr NMJs unload similar levels of FM1-­‐43 following a one-­‐minute incubation with high K+ solution. Mean fluorescence intensity as a percentage of loading fluorescence ± SEM: control = 48.04 ± 4.16, n = 9; gyr = 55.64 ± 5.58, n = 9. Scale bar = 14 μm. 163 164 Figure 7. Ten minutes after incubation with a high-­‐potassium solution, gyr synaptic vesicle size is more homogeneous. A) Same as Figure 4C. At rest, gyr synaptic vesicles are larger and more variable in size. B) Following a five-­‐minute potassium shock and ten-­‐minute rest period, the distribution of synaptic vesicle diameter in gyr mutants is much more similar to controls. C) Scatter plots of the average synaptic vesicle diameter (excluding vesicles >60 nm) of each bouton included in the analysis. Only boutons in which more than 30 synaptic vesicles could be measured were used in the analysis. Mean synaptic vesicle diameter following the potassium shock is not significantly different in gyr mutants relative to controls (p = 0.89, Student’s t-­‐test). Average synaptic vesicle diameter (in nm) ± SEM: control = 42.95 ± 0.63, n = 19; gyr = 43.06 ± 0.48, n = 22. D) A cumulative frequency diagram illustrates the differences in vesicle diameter before (Pre-­‐K) and after (Post-­‐K) potassium shock. The vertical lines mark the region between 30 and 60 nm, which is displayed in part (E) to emphasize the region of the graph that includes synaptic vesicles. All synaptic vesicles and cisternae (up to 150 nm) that could be measured are represented in this graph, although some data points lie beyond 150 nm (especially in the Post-­‐K condition), and therefore the cumulative frequency does not reach 1.0. E) The same data presented in (D) including only the values between 30 nm and 60 nm following normalization. Due to the increased number of cisternae in both control and gyr boutons following the potassium shock, only ~80% of the measured structures lie in the region from 30 to 60 nm, as compared to controls at rest where ~97% of structures are in this range. Therefore, we normalized the data within this region to more accurately compare the distribution of vesicles 30 to 60 nm in diameter. To normalize the data, we recalculated the cumulative frequency distributions of each bouton using only the structures less than 60 nm in diameter. These cumulative frequency distributions were then averaged together to give the overall distribution. Prior to the potassium shock, the gyr distribution is shifted to the right compared to the control. Following the potassium shock, gyr and control synaptic vesicles have much more similar distributions. 165 Figure 8. mEJP and EJP amplitude is normal in gyr mutants. A) Although synaptic vesicle diameter is increased in gyr larvae, mEJP amplitude is not significantly different between gyr and control (p = 0.71, Student’s t-­‐test). Several representative traces from control and gyr larvae are depicted. Average mEJP amplitude (in mV) ± SEM: control = 1.23 ± 0.07, n = 6; gyr = 1.18 ± 0.11, n = 7. B) Evoked release also is not significantly different between gyr and control larvae (p = 0.63, Student’s t-­‐test). Average EJP amplitude (in mV) ± SEM: control = 11.8 ± 1.87, n = 9; gyr = 13.0 ± 1.40, n = 8. 166 167 Figure 9. gyr mutants display increased facilitation and have more asynchronous release at higher stimulation frequencies. A) At a calcium concentration of 0.2 mM, gyr mutants have an increase in the paired-­‐pulse ratio (PPR) compared to controls at a variable interstimulus time (VIST) of 50 ms but not 100 ms. The PPR is defined as the amplitude of the second peak divided by the first peak. B) At 10 Hz and 20 Hz stimulation frequencies, the gyr mutant displays an increase in facilitation that is further enhanced at 20 Hz. The bar graphs indicate the average facilitation at the beginning of stimulation (Initial Phase) and at the end of 500 stimuli (Late Phase). While gyr (red) has higher levels of facilitation compared to controls (black) at both 10 and 20 Hz, this difference is more pronounced at 20 Hz. C) When stimulated at high frequency for an extended period of time (1,500 stimuli at 20 Hz), gyr larvae display an increase in facilitation that persists for approximately 500 stimuli and then declines to levels similar to controls (see [D]). Presynaptic expression of synaptogyrin rescues the phenotype (elavc155-­GAL4; gyr1; UAS-­myc-­gyrin/+). D) For each stimulus from the traces in (C), we plotted the peak current (EJC amplitude) and the amount of charge transferred (total current) separately. While both plots depict an increase in gyr relative to controls during the early phase of stimulation, the difference in the amount of charge transferred between gyr and control is greater than the difference in peak amplitude. E) The same data from (C) and (D) are plotted to compare the amount of charge transferred by a stimulus relative to its peak amplitude. In control and gyr larvae, the amount of current transferred increases as the EJC amplitude increases. However, compared to controls, stimuli from gyr animals transfer a greater amount of charge at similar amplitudes. This suggests that synaptic vesicle release is more asynchronous in gyr mutants. When traces from the 20 Hz stimulation protocol are normalized (lower panel), the gyr peak is broader than the control peak, indicating that there is a component of release that is delayed in the mutant larvae. F) Due to the increased amount of charge transferred by EJCs in gyr larvae, the cumulative release is also enhanced. After 1,500 stimuli, gyr larvae release a total of ~600 nC of charge, while controls release ~375 nC, which is about 40% less total charge over the course of stimulation. However, as stimulation proceeds, a steady state is reached in which gyr and controls transfer charge at a similar rate. This is indicated by the gray line, which represents the difference in release between gyr and control. This line plateaus around 150 nC, indicating the point after which gyr and control larvae begin to transfer charge at a more similar rate. 168 Chapter 4 Conclusions and Perspectives Robin J. Stevens The Picower Institute for Learning and Memory, Department of Biology, Massachusetts Institute of Technology, Cambridge, MA 02139 169 Summary Although synaptogyrin and synaptophysin were identified as integral synaptic vesicle proteins decades ago, their exact role in the synaptic vesicle cycle remains obscure. We have attempted to elucidate the function of synaptogyrin by generating a null mutant in Drosophila, which has only a single gyrin isoform and no physin homolog, thereby allowing us to avoid issues of redundancy. Drosophila synaptogyrin is abundantly expressed in neurons and enriched in synaptic vesicles, which suggests that it is likely a true synaptogyrin homolog. Initial characterization of a gyr mutant revealed no measurable changes in viability, gross synaptic morphology, basal neuro-­‐
transmission, or male courtship behavior. However, in-­‐depth analysis revealed defects in the regulation of synaptic vesicle diameter and in the recovery of synaptic vesicles from endocytic cisternae, as well as alterations in synaptic facilitation. These results suggest that synaptogyrin, while not required for neurotransmission, does modulate synaptic vesicle exo-­‐endocytosis, especially during high-­‐intensity stimulation. The loss of synaptogyrin and synaptophysin results in relatively mild phenotypes While we were able to find several alterations in synaptic vesicle recycling and neurotransmission in gyr mutants, it is intriguing that the loss of a highly conserved synaptic vesicle protein does not result in more dramatic phenotypes. Mouse and nematode synaptogyrin/synaptophysin double knockouts also have relatively mild phenotypes, suggesting that this observation is not unique to Drosophila (Janz et al., 1999; Abraham et al., 2006; Abraham et al., 2011). Synaptogyrin and synaptophysin are not essential for synaptic vesicle exocytosis, endocytosis, or basic behaviors. However, this does not necessarily mean that these proteins do not impact 170 neurotransmission in a manner that is evolutionarily relevant. Indeed, several lines of evidence suggest that synaptophysin and synaptogyrin promote the efficient recycling of synaptic vesicles. For example, an increase in clathrin-­‐coated vesicles was seen in the C. elegans triple knockout (Abraham et al., 2006), the mouse synaptophysin knockout (Spiwoks-­‐Becker et al., 2001), as well as in the squid giant axon when the synaptophysin-­‐dynamin interaction was inhibited (Daly et al., 2000). These observations raise the possibility that synaptophysin and/or synaptogyrin are involved in a clathrin-­‐independent version of endocytosis and that clathrin-­‐mediated endo-­‐
cytosis is upregulated in their absence to counteract for the loss of the synaptogyrin/ synaptophysin-­‐mediated endocytic pathway. Under relatively mild stimulation conditions, it appears as though clathrin-­‐mediated endocytosis is able to compensate for the loss of synaptophysin and/or synaptogyrin. However, under more intense stimulation, the clathrin-­‐mediated endocytic pathway alone is insufficient and alterations in synaptic function begin to emerge. While this hypothesis is intriguing, the rapid inactivation of clathrin in Drosophila through the use of the FlAsH-­‐FALI technique results in a complete block in synaptic vesicle reformation, suggesting that all synaptic vesicles are formed through clathrin-­‐dependent mechanisms at the Drosophila larval NMJ (Heerssen et al., 2008). Studies in mice have established that synaptophysin and synaptogyrin impact synaptic plasticity as well as certain aspects of learning and memory (Janz et al., 1999; Schmitt et al., 2009). While we were unable to identify learning and memory deficits in the Drosophila gyr mutant, it certainly is possible that the loss of synaptogyrin does have important evolutionarily relevant behavioral consequences that we were simply 171 incapable of isolating in our initial examination. For example, synaptogyrin may be required during periods of intense motor activity such as flight and predator avoidance, or it may impact visual acuity, olfaction, temperature tolerance, or social interactions. Behavioral experiments in a controlled laboratory setting cannot fully replicate the unique stresses of life in the natural environment, and it is therefore difficult to determine exactly how synaptogyrin may impact the fitness of an organism in the wild. Interestingly, several recent studies have associated mutations in the synaptogyrin locus with schizophrenia susceptibility in humans (Verma et al., 2004; Verma et al., 2005; Cheng and Chen, 2007; Iatropoulos et al., 2009). It may be the case that the very same alterations in synaptic plasticity and synaptic vesicle recycling observed in model organism knockouts also may have the ability to impact human thought and behavior. Phylogenetic analysis of synaptophysin and synaptogyrin When we searched for the evolutionary emergence of gyrins, physins, and other MARVEL proteins, we were surprised to find synaptogyrin and synaptophysin homologs in organisms without nervous systems. The ancestral functions of physins and gyrins are unknown, although one likely possibility is that their MARVEL transmembrane regions were used in some aspect of membrane organization. For example, other MARVEL proteins such as MAL and MAL2 mediate trafficking of proteins and lipids to the apical membrane (Puertollano and Alonso, 1999; de Marco et al., 2002), while occludin localizes to tight junctions (Furuse et al., 1993). Many MARVEL domain-­‐containing proteins are associated with lipid rafts, which suggests that the MARVEL domain itself has the ability to form membrane microdomains enriched in lipids and proteins that mediate processes such as cellular adhesion, vesicle biogenesis, 172 intracellular trafficking, or intercellular signaling (Thiele et al., 2000; de Marco et al., 2001; Sanchez-­‐Pulido et al., 2002; Raleigh et al., 2010). The ancestral gyrins and physins may have performed similar roles in early metazoans or in their unicellular ancestors, and more recently in evolution they were co-­‐opted by the nervous system to perform specific functions related to synaptic vesicle exo-­‐endocytosis. The presence of two distinct physin and gyrin families in vertebrates with multiple paralogs suggests that several rounds of gene duplication occurred after the split between the ancestral gyrin and the ancestral physin (Hubner et al., 2002). It is currently unknown whether the ancestral physins and gyrins were more similar to the ubiquitous pantophysin and cellugyrin or if they more closely resembled neuronal-­‐specific synaptogyrin and synaptophysin. The absence of a Drosophila synaptophysin homolog and an N. vectensis synaptogyrin homolog raises the possibility that the functions of physins and gyrins are interchangeable. Indeed, the observation that synaptic plasticity defects are enhanced in the mouse double knockout strongly supports the idea that synaptogyrin and synaptophysin are to some extent functionally redundant (Janz et al., 1999). However, it seems unlikely that both proteins would be so highly conserved across evolution if they did not serve important and unique functions in the organisms that have both a synaptogyrin and a synaptophysin homolog. Due to the absence of other physin and gyrin isoforms and the relative ease of transgene expression in Drosophila, the gyr mutant provides an opportunity to examine the relative contributions of gyrins and physins to different aspects of the exo-­‐endocytic cycle. In the future, we would like to perform rescue experiments with the mammalian synaptogyrin and synaptophysin 173 isoforms to determine if any (or all) of the phenotypes we identified in gyr mutants can be rescued by mammalian synaptophysin or synaptogyrin. These experiments could provide insight into whether Drosophila synaptogyrin performs a role more similar to mammalian synaptogyrin or synaptophysin as well as information about the level of functional conservation between mammalian synaptophysin and synaptogyrin. Structure-­function analysis of synaptogyrin and synaptophysin Relatively little is known about how synaptophysin and synaptogyrin function is regulated at the synapse. Mammalian synaptogyrin and synaptophysin contain several tyrosine residues in their C-­‐termini that are phosphorylated by the tyrosine kinases pp60c-­‐src and c-­‐fyn (Sudhof et al., 1987; Barnekow et al., 1990; Linstedt et al., 1992; Janz and Sudhof, 1998; Janz et al., 1999), and synaptophysin appears to be a calcium-­‐
dependent substrate of CaM kinase II (Rubenstein et al., 1993). Interestingly, synaptophysin tyrosine phosphorylation increases in parallel with changes in synaptic plasticity, as indicated by higher levels of synaptophysin phosphorylation following LTP induction in hippocampal slices (Mullany and Lynch, 1998). However, exactly which residues of mammalian synaptogyrin and synaptophysin undergo phosphorylation and the functional consequences of protein modification are currently unknown. Part of the difficulty in isolating the targets of phosphorylation in mammalian synaptophysin has been due to the lack of trypsin cleavage sites in the protein’s C-­‐terminus, which hinders the identification of phosphorylated residues via mass spectroscopy (Evans and Cousin, 2005). We have not yet determined whether Drosophila synaptogyrin is a target of phosphorylation, although this protein does contain several cytoplasmic tyrosine and 174 serine residues. However, since the N-­‐ and C-­‐termini are among the least conserved regions of physins and gyrins, it is difficult to identify evolutionarily conserved phosphorylation sites (Hubner et al., 2002). Nevertheless, it would be informative to perform in vitro kinase assays on Drosophila synaptogyrin to determine if the protein can undergo phosphorylation. If so, identification of phosphorylated residues followed by generation of phosphomimetic and phospho-­‐incompetent versions of Drosophila synaptogyrin for use in rescue experiments may provide critical insight into the protein’s function and regulation. If mammalian synaptophysin and/or synaptogyrin are capable of rescuing some (or all) of the gyr phenotypes, additional rescue experiments could be performed with phosphomimetic and phospho-­‐incompetent versions of these proteins as well. However, since mammalian synaptophysin contains nine tyrosine and four serine residues on its C-­‐terminus alone, this may be an ambitious undertaking (Evans and Cousin, 2005). Additional information about synaptogyrin function in Drosophila could come from deletion and mutagenesis studies to determine the importance of various protein domains in synaptogyrin’s role at the synapse. Since the C-­‐termini of mammalian synaptophysin and synaptogyrin are known to interact with dynamin in a calcium-­‐
dependent manner (Daly and Ziff, 2002), it would be informative to determine if the loss of this region impacts synaptic vesicle recycling in Drosophila. However, the C-­‐
terminus of synaptogyrin also has been identified as a critical region for protein targeting to the synaptic terminal in C. elegans (Zhao and Nonet, 2001). Therefore, it would be difficult to examine the influence of the Drosophila synaptogyrin C-­‐terminus 175 on endocytosis through deletion studies if the loss of this region results in protein mislocalization or degradation. In-­‐depth structure-­‐function analysis of the synaptogyrin paralog cellugyrin demonstrated that vesicle biogenesis in PC12 cells was greatly influenced by two short hydrophobic regions, one located in the lumen between the first two transmembrane domains and the other found in the cytoplasm directly after the fourth transmembrane domain (Belfort et al., 2005). Furthermore, the length of the short cytoplasmic loop (between the second and third transmembrane helices) also significantly impacted cellugyrin function. Taken together, these observations led to a hypothesis proposing that cellugyrin forms a conical structure with the luminal ends of the transmembrane helices situated more closely together than the cytoplasmic ends, thereby directly imparting curvature on membranes (Belfort et al., 2005). Given that the lengths of the regions between the transmembrane helices are conserved among the gyrin paralogs, it is possible that synaptogyrin’s transmembrane domains may also directly promote membrane curvature and thereby impact vesicle formation (Hubner et al., 2002). Thus it would be worthwhile to perform similar structure-­‐function analysis of Drosophila synaptogyrin to determine if the aforementioned regions of cellugyrin are also essential for synaptogyrin function. The role of synaptogyrin and synaptophysin in endocytosis Synaptic vesicle biogenesis, whether from an endosomal compartment or the plasma membrane, requires the recruitment of a wide variety of essential synaptic vesicle proteins (e.g., synaptotagmin and synaptobrevin) and the preferential exclusion of certain proteins such as the t-­‐SNARE syntaxin. Maintaining the correct complement 176 of synaptic vesicle proteins could theoretically be achieved through a “kiss-­‐and-­‐run” endocytic mechanism, which would allow a transiently fused vesicle to directly reform following fusion pore closure rather than fully collapsing into the plasma membrane. While rapid endocytosis has been observed at certain synapses such as the calyx of Held (Wu et al., 2005), the contribution of kiss-­‐and-­‐run endocytosis at the Drosophila NMJ is highly controversial (Verstreken et al., 2002; Dickman et al., 2005; He and Wu, 2007). An alternative way to promote proper synaptic vesicle sorting would be through the creation of microdomains on the plasma membrane or endosomes that are enriched in the lipids and proteins found on synaptic vesicles. Synaptophysin, through its interactions with cholesterol and synaptobrevin, in addition to its ability to multimerize, is an intriguing candidate to seed this sort of microdomain to facilitate vesicle biogenesis (Rehm et al., 1986; Thomas et al., 1988; Calakos and Scheller, 1994; Washbourne et al., 1995; Thiele et al., 2000). Whether synaptogyrin could perform a similar function is unclear because an interaction between synaptogyrin and cholesterol was not tested (Thiele et al., 2000), and mammalian synaptogyrin and synaptobrevin do not appear to interact directly (Edelmann et al., 1995). While we have not yet investigated the lipid-­‐binding capabilities of Drosophila synaptogyrin, preliminary results of a GST pull-­‐down experiment using Drosophila synaptobrevin identified a potential interaction with synaptogyrin (data not shown). However, this result will need to be verified to ensure that this protein interaction is specific and not merely a consequence of the two proteins colocalizing to synaptic vesicles. Synaptophysin also has been suggested to promote endocytosis via a calcium-­‐
dependent interaction with dynamin, and it may recruit dynamin to sites of synaptic 177 vesicle exocytosis (Daly and Ziff, 2002). Like synaptophysin, mammalian synaptogyrin appears to bind dynamin in a calcium-­‐dependent manner, and therefore may contribute to the recruitment of endocytic machinery; however, the synaptogyrin-­‐dynamin association is not as strong as the synaptophysin-­‐dynamin interaction (Daly and Ziff, 2002). We have conducted preliminary experiments to investigate whether Drosophila synaptogyrin is capable of binding dynamin in a calcium-­‐dependent manner. We performed a GST pull-­‐down experiment using the C-­‐terminus of synaptogyrin – the region of synaptophysin and synaptogyrin that binds dynamin in mammals (Daly and Ziff, 2002). Preliminary results indicate that Drosophila synaptogyrin and dynamin do not significantly interact in either the presence or absence of calcium (data not shown), suggesting that this interaction is not conserved in Drosophila. Therefore, it remains unclear whether Drosophila synaptogyrin can promote exo-­‐endocytosis through the same protein interactions observed with mammalian synaptophysin or synaptogyrin. Further analysis of synaptogyrin’s role in vesicle biogenesis and endocytosis We were intrigued to discover that the loss of synaptogyrin results in an increase in endocytic cisternae ten minutes after intense stimulation with high potassium. While we put forward a model in which synaptogyrin promotes the budding of synaptic vesicles from endosomal-­‐like compartments (see Chapter 3, Discussion), we currently cannot rule out an alternative explanation, namely that synaptogyrin normally inhibits homotypic fusion of synaptic vesicles or that it prevents the fusion of synaptic vesicles with endosomal compartments after periods of extreme neuronal activity. If this were the case, the loss of synaptogyrin would enhance the frequency of 178 this type of fusion, which could account for the increase in cisternae we observed ten minutes after potassium stimulation. Indeed, the small GTPase Rab5, which regulates endocytic trafficking to early endosomes, has been identified as an inhibitor of homotypic fusion, indicating that there are mechanisms in place at the presynaptic terminal to prevent synaptic vesicles from fusing with one another (Bucci et al., 1992; Shimizu et al., 2003). Mutations in Rab5 lead to an increase in homotypic fusion and consequently result in significant misregulation of synaptic vesicle size, and this effect is exacerbated when synaptic activity is low (Shimizu et al., 2003). Furthermore, there also is evidence suggesting that synaptic vesicles can fuse with endocytic cisternae at Drosophila larval NMJs under intense stimulation conditions. When FM1-­‐43 dye was loaded for 1.5 minutes and subsequently photocoverted to create an electron-­‐dense product visible by electron microscopy, a subset of cisternae had irregular membrane labeling, as if loaded and unloaded vesicles had fused together (Akbergenova and Bykhovskaia, 2009). Furthermore, labeled synaptic vesicles were occasionally observed in extremely close proximity to unlabeled cisternae, potentially indicating fusion between these structures. Thus, the possibility that synaptogyrin may inhibit homotypic fusion is worthy of further investigation. To distinguish between the two possibilities that synaptogyrin either promotes synaptic vesicle biogenesis from endocytic cisternae or inhibits synaptic vesicle fusion with endosomal-­‐like compartments, we could repeat the previously described FM1-­‐43 experiment in gyr mutants. If synaptogyrin merely regulates synaptic vesicle budding from cisternae, we would expect that the extent of irregular membrane labeling would 179 not be enhanced in the gyr mutant background. However, if synaptogyrin inhibits homotypic fusion, there should be an increase in the degree of uneven cisternal labeling in gyr boutons. Examination of the genetic interaction between Rab5 and synaptogyrin may provide further insights into whether these proteins function in a similar pathway. Furthermore, if synaptogyrin promotes vesicle biogenesis from endocytic cisternae, we would predict that gyr mutants would retain a greater amount of FM1-­‐43 dye in cisternae following potassium-­‐induced uptake via bulk endocytosis. Our initial FM1-­‐43 experiments did not reveal a significant difference in dye loading or unloading in gyr mutants relative to controls. However, our unloading experiments were performed with non-­‐physiological high-­‐potassium saline, which may induce exocytosis of cisternae in addition to synaptic vesicles. This may have prevented us from identifying a delay in the movement of FM1-­‐43 from cisternae to synaptic vesicles. If we were to unload the FM1-­‐43 dye with more physiologically relevant stimuli (i.e., with nerve stimulation), we would expect to see greater retention of FM1-­‐43 in gyr boutons, as more of the dye would be trapped in cisternae compared to controls. Other stimulation protocols, including those that selectively label the recycling synaptic vesicle pool or the reserve pool, may provide additional insight into how the loss of synaptogyrin affects synaptic vesicle recycling. Examination of genetic interactions with Drosophila synaptogyrin Previous studies in Drosophila have examined many of the critical proteins involved in regulating clathrin-­‐mediated endocytosis at the larval NMJ. Comparing the endocytic phenotypes caused by mutations in these genes with those observed in our gyr mutant may provide further insight into synaptogyrin function. Intriguingly, 180 several of these endocytic mutants share phenotypes in common with gyr animals. For example, the loss of synaptojanin, a polyphosphoinositide phosphatase that participates in the uncoating of the clathrin coat (Cremona et al., 1999), or AP-­‐180/lap (lap is the Drosophila homolog), a clathrin adaptor that promotes clathrin assembly (Ahle and Ungewickell, 1986), leads to a decrease in synaptic vesicle density and a dramatic increase in synaptic vesicle size, which we also observed in a subset of gyr boutons (Zhang et al., 1998a; Verstreken et al., 2003; Dickman et al., 2005). The loss of endophilin, a protein that binds dynamin and recruits synaptojanin to sites of endocytosis, results in many of the same phenotypes as those seen in the synaptojanin mutant, suggesting it may also play a role in removing clathrin coats (Ringstad et al., 1997; Verstreken et al., 2002; Verstreken et al., 2003; Dickman et al., 2005). However, disruption of synaptojanin, AP-­‐180/lap, or endophilin function also resulted in reduced uptake of FM1-­‐43, indicating a general defect in endocytosis that we did not observe in gyr larvae (Verstreken et al., 2002; Verstreken et al., 2003; Dickman et al., 2005). Under basal conditions, endophilin and lap mutants have an increase in mEJP amplitude, while synaptojanin mutants display an increase in quantal amplitude following a tetanus (Zhang et al., 1998a; Verstreken et al., 2002; Dickman et al., 2005). Although we did not examine mEJP amplitude after intense stimulation, gyr larvae did not have a significant increase in quantal size under basal conditions. It should be noted that, while gyr larvae have a subset of boutons with dramatically enhanced synaptic vesicle diameter, many boutons are comparable to controls in terms of synaptic vesicle size and density, which may explain why we did not observe a measurable increase in mEJP amplitude, especially if the more abnormal gyr boutons have a lower level of synaptic activity. 181 Furthermore, the synaptic vesicles near active zones in gyr larvae were normal in size, while the other endocytic mutants often had large synaptic vesicles in close proximity to release sites (Zhang et al., 1998a; Verstreken et al., 2002; Dickman et al., 2005). Overall, the endocytic phenotypes observed in synaptojanin, endophilin, and lap mutants are much greater in severity than those in gyr animals. While gyr larvae also have an increase in synaptic vesicle size, they do not have defects in FM1-­‐43 uptake, a dramatic decrease in synaptic vesicle density, or alterations in quantal size. Therefore, it is difficult to determine where synaptogyrin might act in the clathrin-­‐mediated endocytic pathway based on phenotypes alone. Synaptogyrin likely does not play an essential role in clathrin coat formation, as the inactivation of clathrin via FlAsH-­‐FALI results in the complete loss of vesicles (Heerssen et al., 2008). Similarly, a severe loss-­‐
of-­‐function mutation in the α-­adaptin gene, a component of the AP-­‐2 complex, results in embryonic lethality and the absence of vesicles and collared pits at synapses (Gonzalez-­‐
Gaitan and Jackle, 1997). Synaptogyrin is also not necessary for synaptic vesicle fission, since the inactivation of dynamin using the shibire temperature-­‐sensitive mutation also completely blocks vesicle formation and induces the formation of membrane invaginations and collared pits (Koenig and Ikeda, 1983, 1989). Instead, synaptogyrin likely plays a modulatory or regulatory role in synaptic vesicle formation. Alternatively, it may only promote synaptic vesicle endocytosis under certain conditions or function in a subset of endocytic pathways. Interestingly, a recent study investigating a C. elegans synaptogyrin mutant indentified synthetic phenotypes with synaptojanin and endophilin mutants (Abraham et al., 2011). The observation that the loss of synaptogyrin enhances the phenotypes of known regulators of clathrin-­‐mediated endocytosis promotes the 182 previously discussed hypothesis that synaptogyrin may participate in an alternative clathrin-­‐independent recycling pathway. However, synaptogyrin may instead promote efficient clathrin-­‐mediated endocytosis through a mechanism that parallels the function of synaptojanin and/or endophilin. We would like to repeat these genetic interaction experiments in Drosophila by combining the gyr mutation with mutations in genes such as endophilin, lap, and synaptojanin to gain further insight into how synaptogyrin function impacts synaptic vesicle endocytosis. We also would like to determine the extent to which Drosophila synaptogyrin promotes synaptic vesicle endocytosis from the plasma membrane as opposed to endosomes or cisternae. To accomplish this, we could inhibit bulk endocytosis in our gyr mutant using a temperature-­‐sensitive allele of rolling blackout (rbots), a Drosophila EFR3 integral membrane lipase (Huang et al., 2006; Vijayakrishnan and Broadie, 2006; Vijayakrishnan et al., 2009). At the restrictive temperature, rbots mutants fail to form endocytic cisternae in response to incubation with high-­‐potassium saline but are still capable of generating synaptic vesicles, presumably by endocytosis from the plasma membrane (Vijayakrishnan et al., 2009). If Drosophila synaptogyrin solely participates in synaptic vesicle budding from cisternae, the gyr mutation should not enhance the rbots phenotype. However, if gyr also mediates synaptic vesicle endocytosis from the plasma membrane, a gyr/rbots double mutant would be expected to have endocytic deficits beyond those seen in the rbots mutant alone. Further analysis of the genetic interactions of gyr in combination with additional mutations known to affect endocytosis and exocytosis may provide added insight into the role of synaptogyrin in the synaptic vesicle cycle. 183 Conclusions While we have provided additional evidence implicating synaptogyrin in the regulation of synaptic vesicle exo-­‐endocytosis, many questions still remain. Although we identified alterations in synaptic vesicle recycling dynamics and synaptic plasticity in gyr larvae, we still do not know exactly how synaptogyrin impacts synaptic vesicle fusion and biogenesis. Further studies examining the protein-­‐binding and genetic interactions of synaptogyrin, along with comprehensive structure-­‐function analysis may help illuminate the mechanisms behind synaptogyrin function at the synapse. Finally, we would like to determine how (or if) the loss of synaptogyrin impacts Drosophila on a behavioral level and whether the observed changes in synaptic function lead to deficits in learning and memory. 184 References Abraham C, Bai L, Leube RE (2011) Synaptogyrin-­‐dependent modulation of synaptic neurotransmission in Caenorhabditis elegans. Neuroscience 190:75-­‐88. Abraham C, Hutter H, Palfreyman MT, Spatkowski G, Weimer RM, Windoffer R, Jorgensen EM, Leube RE (2006) Synaptic tetraspan vesicle membrane proteins are conserved but not needed for synaptogenesis and neuronal function in Caenorhabditis elegans. Proc Natl Acad Sci U S A 103:8227-­‐8232. Ahle S, Ungewickell E (1986) Purification and properties of a new clathrin assembly protein. EMBO J 5:3143-­‐3149. Akbergenova Y, Bykhovskaia M (2009) Enhancement of the endosomal endocytic pathway increases quantal size. Mol Cell Neurosci 40:199-­‐206. Barnekow A, Jahn R, Schartl M (1990) Synaptophysin: a substrate for the protein tyrosine kinase pp60c-­‐src in intact synaptic vesicles. Oncogene 5:1019-­‐1024. Belfort GM, Bakirtzi K, Kandror KV (2005) Cellugyrin induces biogenesis of synaptic-­‐
like microvesicles in PC12 cells. J Biol Chem 280:7262-­‐7272. Bucci C, Parton RG, Mather IH, Stunnenberg H, Simons K, Hoflack B, Zerial M (1992) The small GTPase rab5 functions as a regulatory factor in the early endocytic pathway. Cell 70:715-­‐728. Calakos N, Scheller RH (1994) Vesicle-­‐associated membrane protein and synaptophysin are associated on the synaptic vesicle. J Biol Chem 269:24534-­‐24537. Cheng MC, Chen CH (2007) Identification of rare mutations of synaptogyrin 1 gene in patients with schizophrenia. J Psychiatr Res 41:1027-­‐1031. Cremona O, Di Paolo G, Wenk MR, Luthi A, Kim WT, Takei K, Daniell L, Nemoto Y, Shears SB, Flavell RA, McCormick DA, De Camilli P (1999) Essential role of phosphoinositide metabolism in synaptic vesicle recycling. Cell 99:179-­‐188. Daly C, Ziff EB (2002) Ca2+-­‐dependent formation of a dynamin-­‐synaptophysin complex: potential role in synaptic vesicle endocytosis. J Biol Chem 277:9010-­‐9015. Daly C, Sugimori M, Moreira JE, Ziff EB, Llinas R (2000) Synaptophysin regulates clathrin-­‐independent endocytosis of synaptic vesicles. Proc Natl Acad Sci U S A 97:6120-­‐6125. de Marco MC, Kremer L, Albar JP, Martinez-­‐Menarguez JA, Ballesta J, Garcia-­‐Lopez MA, Marazuela M, Puertollano R, Alonso MA (2001) BENE, a novel raft-­‐associated protein of the MAL proteolipid family, interacts with caveolin-­‐1 in human endothelial-­‐like ECV304 cells. J Biol Chem 276:23009-­‐23017. de Marco MC, Martin-­‐Belmonte F, Kremer L, Albar JP, Correas I, Vaerman JP, Marazuela M, Byrne JA, Alonso MA (2002) MAL2, a novel raft protein of the MAL family, is an essential component of the machinery for transcytosis in hepatoma HepG2 cells. J Cell Biol 159:37-­‐44. Dickman DK, Horne JA, Meinertzhagen IA, Schwarz TL (2005) A slowed classical pathway rather than kiss-­‐and-­‐run mediates endocytosis at synapses lacking synaptojanin and endophilin. Cell 123:521-­‐533. Edelmann L, Hanson PI, Chapman ER, Jahn R (1995) Synaptobrevin binding to synaptophysin: a potential mechanism for controlling the exocytotic fusion machine. EMBO J 14:224-­‐231. 185 Evans GJ, Cousin MA (2005) Tyrosine phosphorylation of synaptophysin in synaptic vesicle recycling. Biochem Soc Trans 33:1350-­‐1353. Furuse M, Hirase T, Itoh M, Nagafuchi A, Yonemura S, Tsukita S (1993) Occludin: a novel integral membrane protein localizing at tight junctions. J Cell Biol 123:1777-­‐
1788. Gonzalez-­‐Gaitan M, Jackle H (1997) Role of Drosophila alpha-­‐adaptin in presynaptic vesicle recycling. Cell 88:767-­‐776. He L, Wu LG (2007) The debate on the kiss-­‐and-­‐run fusion at synapses. Trends Neurosci 30:447-­‐455. Heerssen H, Fetter RD, Davis GW (2008) Clathrin dependence of synaptic-­‐vesicle formation at the Drosophila neuromuscular junction. Curr Biol 18:401-­‐409. Huang FD, Woodruff E, Mohrmann R, Broadie K (2006) Rolling blackout is required for synaptic vesicle exocytosis. J Neurosci 26:2369-­‐2379. Hubner K, Windoffer R, Hutter H, Leube RE (2002) Tetraspan vesicle membrane proteins: synthesis, subcellular localization, and functional properties. Int Rev Cytol 214:103-­‐159. Iatropoulos P, Gardella R, Valsecchi P, Magri C, Ratti C, Podavini D, Rossi G, Gennarelli M, Sacchetti E, Barlati S (2009) Association study and mutational screening of SYNGR1 as a candidate susceptibility gene for schizophrenia. Psychiatr Genet 19:237-­‐243. Janz R, Sudhof TC (1998) Cellugyrin, a novel ubiquitous form of synaptogyrin that is phosphorylated by pp60c-­‐src. J Biol Chem 273:2851-­‐2857. Janz R, Sudhof TC, Hammer RE, Unni V, Siegelbaum SA, Bolshakov VY (1999) Essential roles in synaptic plasticity for synaptogyrin I and synaptophysin I. Neuron 24:687-­‐700. Koenig JH, Ikeda K (1983) Evidence for a presynaptic blockage of transmission in a temperature-­‐sensitive mutant of Drosophila. J Neurobiol 14:411-­‐419. Koenig JH, Ikeda K (1989) Disappearance and reformation of synaptic vesicle membrane upon transmitter release observed under reversible blockage of membrane retrieval. J Neurosci 9:3844-­‐3860. Linstedt AD, Vetter ML, Bishop JM, Kelly RB (1992) Specific association of the proto-­‐
oncogene product pp60c-­‐src with an intracellular organelle, the PC12 synaptic vesicle. J Cell Biol 117:1077-­‐1084. Mullany PM, Lynch MA (1998) Evidence for a role for synaptophysin in expression of long-­‐term potentiation in rat dentate gyrus. Neuroreport 9:2489-­‐2494. Puertollano R, Alonso MA (1999) MAL, an integral element of the apical sorting machinery, is an itinerant protein that cycles between the trans-­‐Golgi network and the plasma membrane. Mol Biol Cell 10:3435-­‐3447. Raleigh DR, Marchiando AM, Zhang Y, Shen L, Sasaki H, Wang Y, Long M, Turner JR (2010) Tight junction-­‐associated MARVEL proteins marveld3, tricellulin, and occludin have distinct but overlapping functions. Mol Biol Cell 21:1200-­‐1213. Rehm H, Wiedenmann B, Betz H (1986) Molecular characterization of synaptophysin, a major calcium-­‐binding protein of the synaptic vesicle membrane. EMBO J 5:535-­‐
541. 186 Ringstad N, Nemoto Y, De Camilli P (1997) The SH3p4/Sh3p8/SH3p13 protein family: binding partners for synaptojanin and dynamin via a Grb2-­‐like Src homology 3 domain. Proc Natl Acad Sci U S A 94:8569-­‐8574. Rubenstein JL, Greengard P, Czernik AJ (1993) Calcium-­‐dependent serine phosphorylation of synaptophysin. Synapse 13:161-­‐172. Sanchez-­‐Pulido L, Martin-­‐Belmonte F, Valencia A, Alonso MA (2002) MARVEL: a conserved domain involved in membrane apposition events. Trends Biochem Sci 27:599-­‐601. Schmitt U, Tanimoto N, Seeliger M, Schaeffel F, Leube RE (2009) Detection of behavioral alterations and learning deficits in mice lacking synaptophysin. Neuroscience 162:234-­‐243. Shimizu H, Kawamura S, Ozaki K (2003) An essential role of Rab5 in uniformity of synaptic vesicle size. J Cell Sci 116:3583-­‐3590. Spiwoks-­‐Becker I, Vollrath L, Seeliger MW, Jaissle G, Eshkind LG, Leube RE (2001) Synaptic vesicle alterations in rod photoreceptors of synaptophysin-­‐deficient mice. Neuroscience 107:127-­‐142. Sudhof TC, Lottspeich F, Greengard P, Mehl E, Jahn R (1987) A synaptic vesicle protein with a novel cytoplasmic domain and four transmembrane regions. Science 238:1142-­‐1144. Thiele C, Hannah MJ, Fahrenholz F, Huttner WB (2000) Cholesterol binds to synaptophysin and is required for biogenesis of synaptic vesicles. Nat Cell Biol 2:42-­‐49. Thomas L, Hartung K, Langosch D, Rehm H, Bamberg E, Franke WW, Betz H (1988) Identification of synaptophysin as a hexameric channel protein of the synaptic vesicle membrane. Science 242:1050-­‐1053. Verma R, Kubendran S, Das SK, Jain S, Brahmachari SK (2005) SYNGR1 is associated with schizophrenia and bipolar disorder in southern India. J Hum Genet 50:635-­‐
640. Verma R, Chauhan C, Saleem Q, Gandhi C, Jain S, Brahmachari SK (2004) A nonsense mutation in the synaptogyrin 1 gene in a family with schizophrenia. Biol Psychiatry 55:196-­‐199. Verstreken P, Kjaerulff O, Lloyd TE, Atkinson R, Zhou Y, Meinertzhagen IA, Bellen HJ (2002) Endophilin mutations block clathrin-­‐mediated endocytosis but not neurotransmitter release. Cell 109:101-­‐112. Verstreken P, Koh TW, Schulze KL, Zhai RG, Hiesinger PR, Zhou Y, Mehta SQ, Cao Y, Roos J, Bellen HJ (2003) Synaptojanin is recruited by endophilin to promote synaptic vesicle uncoating. Neuron 40:733-­‐748. Vijayakrishnan N, Broadie K (2006) Temperature-­‐sensitive paralytic mutants: insights into the synaptic vesicle cycle. Biochem Soc Trans 34:81-­‐87. Vijayakrishnan N, Woodruff EA, 3rd, Broadie K (2009) Rolling blackout is required for bulk endocytosis in non-­‐neuronal cells and neuronal synapses. J Cell Sci 122:114-­‐125. Washbourne P, Schiavo G, Montecucco C (1995) Vesicle-­‐associated membrane protein-­‐
2 (synaptobrevin-­‐2) forms a complex with synaptophysin. Biochem J 305 ( Pt 3):721-­‐724. 187 Wu W, Xu J, Wu XS, Wu LG (2005) Activity-­‐dependent acceleration of endocytosis at a central synapse. J Neurosci 25:11676-­‐11683. Zhang B, Koh YH, Beckstead RB, Budnik V, Ganetzky B, Bellen HJ (1998) Synaptic vesicle size and number are regulated by a clathrin adaptor protein required for endocytosis. Neuron 21:1465-­‐1475. Zhao H, Nonet ML (2001) A conserved mechanism of synaptogyrin localization. Mol Biol Cell 12:2275-­‐2289. 188 
Download